Uploaded by danykan781

Understanding Signal Integrity

advertisement
Understanding Signal Integrity
Library of Congress Cataloging-in-Publication Data
A catalog record for this book is available from the U.S. Library of Congress.
British Library Cataloguing in Publication Data
A catalogue record for this book is available from the British Library.
ISBN-13: 978-1-59693-981-3
Cover design by Vicki Kane
© 2011 Artech House
685 Canton Street
Norwood, MA 02062
All rights reserved. Printed and bound in the United States of America. No part
of this book may be reproduced or utilized in any form or by any means, electronic or mechanical, including photocopying, recording, or by any information
storage and retrieval system, without permission in writing from the publisher.
All terms mentioned in this book that are known to be trademarks or service
marks have been appropriately capitalized. Artech House cannot attest to the
accuracy of this information. Use of a term in this book should not be regarded
as affecting the validity of any trademark or service mark.
10 9 8 7 6 5 4 3 2 1
Understanding Signal Integrity
Stephen C. Thierauf
To Christopher, Kevin, and, as always, Ann
Contents
Preface
xv
CHAPTER 1
Introduction to Signal Integrity
1
1.1 What Is Signal Integrity?
1.2 The Importance of Signal Integrity: The First Transatlantic
Telegraph Cable
1.3 What Is a Pulse?
1.4 Time and Frequency Domains
1.4.1 Line Spectrums
1.4.2 Recreating a Pulse with Sine Waves
1.4.3 Why Does the Frequency Domain Matter to Signal
Integrity Engineers?
1.4.4 Upper Bandwidth
1.5 Power Integrity and Simultaneous Switching Noise
1.5.1 What Is Simultaneous Switching Noise?
1.6 What Is the Dielectric Constant and Loss Tangent?
1.7 Main Points
References
1
1
3
4
4
4
5
7
8
8
9
10
10
CHAPTER 2
The Signal Integrity Process
13
2.1 Introduction
2.2 Why Perform a Signal Integrity Analysis?
2.3 What Is a Typical Signal Integrity Workflow?
2.4 Signal Integrity Worst-Case Analysis
2.4.1 Silicon Worst-Case I/O Models
2.4.2 What Combination of Environmental Effects Are Worst Case?
2.4.3 Using SI Analysis Results During Debug
2.4.4 Electrical Overstresses
2.5 Main Points
References
13
13
14
16
17
17
18
19
19
20
v
vi
Contents
CHAPTER 3
Signal Integrity CAD and Models
21
3.1 Introduction
3.2 I/O Models
3.2.1 What Are Transistor Level Models?
3.2.2 What Are IBIS Models?
3.3 Modeling Transmission Lines
3.4 S-Parameters
3.4.1 What Are S-Parameters?
3.4.2 What Is Insertion Loss?
3.4.3 What Is Reflection Loss?
3.4.4 What Are Touchstone Files?
3.5 Field Solvers
3.5.1 What Are 2D Field Solvers?
3.5.2 What Are 3D Field Solvers?
3.6 Main Points
References
21
21
21
22
23
24
24
24
26
26
26
26
27
28
28
CHAPTER 4
Printed Test and Evaluation Boards
31
4.1 Introduction
4.2 Test Boards
4.2.1 Test Boards as a Chip Debug Platform
4.2.2 Margining Power Supply
4.2.3 Testing Decoupling Capacitor Efficacy
4.2.4 Testing Sensitivity to Clock Quality
4.3 Evaluation Boards
4.4 Roll of the Printed Circuit Layout Shop
4.5 Fabricating and Assembling the Board
4.6 Alternative Methods for Design and Fabrication
4.7 Main Points
References
31
31
32
33
33
33
33
34
35
35
36
36
CHAPTER 5
Printed Circuit Board Construction
37
5.1 Introduction
5.2 How Is a Multilayer Circuit Board Constructed?
5.2.1 How Are Connections Made Between Layers?
5.2.2 What Is the Via Aspect Ratio?
5.3 What Is the Significance of Calling a Circuit Board “FR4”?
5.3.1 Why Is There Such Variability in Dk Between Laminate
Manufacturers?
5.3.2 Why Is the Same Board Produced Differently by Various Fab Shops?
5.3.3 What Are Some Alternatives to FR4?
37
37
37
39
39
40
41
41
Contents
5.4 Circuit Board Traces
5.4.1 Microstrip
5.4.2 Stripline
5.4.3 What Is a Mil and What Does a Trace Thickness in Ounces Mean?
5.4.4 What Copper Types Are Used to Form Traces and What Are Their
Characteristics?
5.4.5 What Is Surface Roughness and What Are Typical Values?
5.4.6 Are Traces Rectangular?
5.5 Main Points
Problems
References
vii
42
43
43
43
45
45
46
46
47
48
CHAPTER 6
Transmission Line Fundamentals
49
6.1 Introduction
6.2 What Is a Transmission Line?
6.2.1 What Is the Signal Return Path?
6.3 Circuit Model of a Transmission Line
6.4 Impedance and Delay
6.4.1 Units and Electrical Length
6.5 How Does a Signal Travel Down the Line?
6.5.1 What Voltage Gets Launched Down a Transmission Line?
6.6 How Does the Current in the Return Path Behave?
6.6.1 How Does Current Flow in Return Paths That Are Not Ground?
6.6.2 What Is the Behavior When One Return Path Is a Related
Power Plane?
6.6.3 How Does the Return Current Flow When the Power Supply
Is Not the Signaling Voltage?
6.7 When Does a Conductor, Via, or Connector Pin Act Like a
Transmission Line?
6.7.1 Why Does the Signal Rise Time Matter But Not the Pulse Width?
6.7.2 Why Does Propagation Delay Time Depend on the Line Length,
But Impedance Does Not?
6.8 Main Points
Problems
References
49
49
49
50
51
52
52
54
55
55
56
57
58
59
59
59
60
61
CHAPTER 7
Understanding Microstrip and Stripline Transmission Lines
63
7.1 Introduction
7.2 DC Resistance and Resistivity
7.3 Stripline and Microstrip Capacitance
7.3.1 What Is the Effective Dielectric Constant?
7.3.2 What Are the Differences Between Microstrip and Stripline
Capacitances?
63
63
64
65
66
viii
Contents
7.3.3 How Does Frequency Alter the Loss Tangent?
7.3.4 What Is Conductance and How Is It Determined?
7.4 Stripline and Microstrip Inductance
7.5 How Does the Skin Effect Change the Trace’s Resistance?
7.5.1 What Is the Proximity Effect?
7.6 How Does the Skin Effect Change the Inductance?
7.7 Understanding Stripline and Microstrip Impedance
7.8 Understanding Stripline and Microstrip Propagation Delay Time
7.8.1 How Do the Dielectric and Effective Dielectric Constants Affect
Delay Time?
7.8.2 How Sensitive Is Delay Time to Changes in the Dielectric Constant?
7.9 Main Points
Problems
References
67
67
68
69
69
71
71
73
73
74
74
76
76
CHAPTER 8
Signal Loss and the Effects of Circuit Board Physical Factors
79
8.1 Introduction
8.2 What Are Transmission Line Loss and Attenuation?
8.2.1 What Effects Do Losses Have on Signals?
8.2.2 How Is Loss Specified?
8.2.3 Converting Between Decibels and Percentage Loss
8.3 What Creates Transmission Line Loss and How Is It Characterized?
8.3.1 Loss Differences Between Microstrip and Stripline Traces
8.4 How Do Impedance and Loop Resistance Affect Conductor Loss?
8.4.1 How Does Surface Roughness Increase Conductor Loss?
8.4.2 Surface Roughness in Perspective
8.4.3 What Can Be Done to Reduce Conductor Losses?
8.5 Understanding Dielectric Losses
8.5.1 Differences in Dielectric Losses Between Stripline and Microstrip
8.5.2 Effects of Temperature and Moisture
8.5.3 What Can Be Done to Reduce Dielectric Losses?
8.6 Summary of Signal Loss and Distortion Characteristics
8.7 Effects of Resin Content and Glass Weave on Delay Time and
Impedance
8.8 Effects of Trace Shape
8.9 Main Points
Problems
References
79
79
79
80
81
81
82
83
84
85
85
86
87
87
88
88
89
89
89
90
91
CHAPTER 9
Understanding Trace-to-Trace Coupling
93
9.1 Introduction
9.2 Understanding Mutual Capacitance and Inductance
9.2.1 What Are the Capacitance and Inductance Matrices?
93
93
93
Contents
ix
9.2.2 Why Do Some Field Solvers Report Capacitances with Negative
Numbers?
9.3 How Does Switching Alter the Trace Inductance and Capacitance?
9.3.1 How Is the Capacitance Affected?
9.3.2 How Is the Inductance Affected?
9.4 What Effect Does Coupling Have on Impedance and Delay?
9.5 What Are Odd and Even Modes?
9.5.1 Can the Odd- and Even-Mode Equations Be Used with More
Than Two Traces?
9.5.2 Why Is the Even- and Odd-Mode Timing the Same for Stripline
But Not Microstrip?
9.5.3 By How Much Does the Even- and Odd-Mode Impedance Change?
9.6 What Are the Circuit Effects When Switching Causes the
Impedance to Change?
9.7 How Is Receiver Timing Affected by In- and Out-of-Phase Switching?
9.8 Main Points
Problems
References
95
95
96
97
98
99
101
101
102
103
103
105
105
106
CHAPTER 10
Understanding Crosstalk
107
10.1 Introduction
10.2 How Is Crosstalk Created and What Are Its Characteristics?
10.3 Far-End Crosstalk
10.3.1 Calculating FEXT
10.3.2 What Are Typical Kf Values?
10.3.3 FEXT Summary
10.4 Near-End Crosstalk
10.4.1 Calculating NEXT
10.4.2 What Are Typical Kb Values?
10.4.3 NEXT Summary
10.5 How Closely Do Calculation and Simulation Agree?
10.5.1 How Well Do 2D Field Solvers Agree?
10.6 Guard Traces
10.6.1 Connecting the Guard Trace
10.7 Main Points
Problems
References
107
107
109
109
111
112
112
112
114
115
115
116
116
117
117
119
121
CHAPTER 11
Understanding Signal Reflections
123
11.1 Introduction
11.2 How Are Reflections Created?
11.2.1 A Physical Analogy for Transmission Line Reflections
11.3 The Reflection Coefficient
123
123
126
126
x
Contents
11.3.1 What Determines the Reflection Coefficient Value?
11.4 How Do the Reflected Waves Combine with the Incident Waves?
11.4.1 Difference Between the Response of a Pulse and a Step
11.5 What Is the Behavior When There Are Multiple Reflections?
11.5.1 What Is the Behavior When There Are Many Pulses?
11.6 Reactive Discontinuities
11.7 Main Points
Problems
References
126
127
127
128
132
132
134
135
135
CHAPTER 12
Termination Strategies
137
12.1 Introduction
12.2 Source Series Termination
12.2.1 Selecting the Resistance Value
12.2.2 Effect of Driver Resistance
12.2.3 Power Savings
12.3 Parallel and Thevenin Termination
12.3.1 Selecting Vtt and Vterm
12.3.2 Thevenin Termination
12.3.3 Connecting Rterm Directly to Vdd or Ground
12.4 Diode Termination
12.4.1 Diode Types
12.5 AC Termination
12.6 Topologies
12.6.1 Single-Load Point-to-Point Termination
12.6.2 Multiple Load Point-to-Point Termination
12.7 Multidrop Lines
12.7.1 Response When Signal Rise Time Is Very Short
12.7.2 Response When Signal Rise Time Is Comparable to the
Transmission Line Delays
12.8 Stubs and Branches
12.8.1 Branches
12.9 Main Points
Problems
References
137
137
137
139
139
139
141
142
143
143
144
145
146
146
148
150
151
152
154
155
157
158
159
CHAPTER 13
Differential Signaling
161
13.1 Introduction
13.2 What Are the Electrical Characteristics of Differential Signaling?
13.2.1 How Is a Differential Pair Observed?
13.2.2 What Is an Eye Diagram?
13.2.3 What Is Intersymbol Interference?
13.2.4 How Are the Effects of Loss Corrected?
161
162
163
163
164
165
Contents
13.3 What Are the Electrical Characteristics of a Differential
Transmission Line?
13.3.1 Why Is the Differential Impedance Twice the Odd-Mode
Impedance?
13.4 How Are Differential Transmission Lines Terminated?
13.4.1 How Should a Diff-Pair Be Terminated When the Propagation
Is Not Fully Odd Mode?
13.4.2 What Prevents Creating Odd-Mode Signals?
13.5 How Are Differential Transmission Lines Created?
13.5.1 Manufacturing Trade-Offs Between Loosely and Tightly
Coupled Pairs
13.5.2 Electrical Trade-Offs Between Loosely and Tightly Coupled Pairs
13.6 What Are Some Suggested Diff-Pair Layout and Routing Rules?
13.6.1 Obtaining the Proper Differential Impedance
13.6.2 Signal Trace Between Diff-Pairs
13.6.3 Keeping Signal Traces Far Away from Diff-Pairs
13.6.4 Avoid Return Path Splits
13.6.5 Route on Same Layers
13.6.6 Minimize Layer Hopping
13.6.7 Use Differential Vias
13.6.8 Match Overall Trace Lengths
13.6.9 Match Lengths on Each Layer
13.6.10 Ensure That Noise Is Equally Coupled to Both Traces
13.7 Main Points
Problems
References
xi
165
166
169
170
172
172
173
174
174
175
176
176
177
177
178
178
178
178
179
181
181
182
CHAPTER 14
Trace and Via Artwork Considerations for Signal Integrity
185
14.1 Introduction
14.2 Mitered Corners
14.3 Routing Near the Board Edge
14.4 Serpentine Traces
14.4.1 How Are Serpentine Delay Lines Modeled?
14.4.2 Effects of Corners
14.4.3 How Long Should the Segments Be?
14.4.4 Guard Traces
14.5 Vias
14.5.1 Via Circuit Model
14.5.2 What Factors Determine Via Capacitance?
14.5.3 What Factors Determine Via Inductance?
14.5.4 What Are the Effects of Different Antipad Diameters?
14.5.5 What Are Nonfunctional Pads?
14.5.6 Can the Via Impedance Be Adjusted?
14.5.7 What Are Differential Vias?
185
185
187
188
190
190
190
191
191
191
192
193
193
193
194
194
xii
Contents
14.6
Main Points
References
195
195
CHAPTER 15
Identifying Common Signal Integrity Problems
199
15.1
15.2
15.3
15.4
15.5
15.6
15.7
15.8
15.9
199
199
201
202
204
205
206
206
206
207
Introduction
Questions to Ask When Reviewing PCB Stackup
Questions to Ask When Reviewing Crosstalk Simulations
Questions to Ask When Reviewing Signal Quality Simulations
Questions to Ask When Reviewing Prelayout Simulations
Questions to Ask When Reviewing Postlayout Simulations
Questions to Ask When Reviewing a Termination Strategy
Questions to Ask When Reviewing PCB Layout Design Rules
Questions to Ask When Reviewing ASIC Driver Selection Choices
References
CHAPTER 16
Solving Common Signal Integrity Problems
209
16.1 Introduction
16.2 Reducing Crosstalk
16.3 Reducing Reflections
16.4 Reducing and Eliminating Simultaneous Switching Noise (SSN)
16.5 Improving Inadequate Timing Margins
16.6 Correcting Intersymbol Interference (ISI)
16.7 Steps to Take When Circuit Board Traces Are Too Lossy
16.8 Options to Reduce Circuit Board Thickness
16.9 Steps to Take When There Are Not Enough Routing Layers
16.10 Steps to Take When a Circuit Board Must Be Cost Reduced
16.11 Steps to Take When Data-Dependent Errors Are Causing System
Failures
References
209
209
210
211
211
212
213
213
214
215
216
216
CHAPTER 17
Calculating Trace and Plane Electrical Values
217
17.1 Introduction
17.2 How to Convert from Decibel Loss to Signal Swing
17.3 Estimating DC Resistance
17.4 Finding Inductance and Capacitance When the Physical Dimensions
Are Not Known
17.5 Finding Stripline Inductance and Capacitance When Impedance
Is Known
17.6 Finding Stripline Inductance and Capacitance When Trace Geometry
Is Known
17.7 Finding Microstrip Inductance and Capacitance When Trace
Geometry Is Known
217
217
218
219
220
220
221
Contents
xiii
17.7.1 Microstrip Inductance
17.7.2 Microstrip Capacitance
221
221
17.8 Estimating Inductance and Capacitance of a Plane
17.8.1 Inductance
17.8.2 Capacitance
17.9 Calculating Trace Loss from a Circuit Model
17.10 Calculating Stripline Loss from Trace Dimensions
17.10.1 Dielectric Loss
17.11 Calculating Microstrip Loss from Trace Dimensions
17.11.1 Dielectric Loss
17.12 Finding Stripline Impedance
17.13 Finding Exposed Microstrip Impedance
17.14 Finding Solder Mask Covered Microstrip Impedance
17.15 Wavelength
References
222
222
223
223
224
224
225
225
226
226
227
228
229
About the Author
231
Index
233
Preface
This is an introductory text for engineers, project leaders, and managers wishing to
understand the fundamentals of signal integrity.
The electrical characteristics of the various types of transmission lines present
on circuit boards are covered in depth, with particular attention to the differences
between them. Also covered in detail are reflections, termination strategies, crosstalk, and information regarding circuit board technology. Signal loss and differential signaling are presented in enough detail to provide the reader with a solid
background in these important topics.
Because the focus is on those new to signal integrity (including managers),
chapters covering the signal integrity process, CAD tools, the value of test and
evaluation circuit boards, and troubleshooting strategies are presented. This material is unique and is intended to provide the newcomer with an overview and the
manager with an understanding of the tasks that signal integrity engineers routinely carry out. The need to perform some of these tasks (such as model validation or
the debugging of individual circuit models) and the uncertainty in the results from
simulation are often unknown to management and to those new to the field and are
not always properly accounted for in project schedules.
Exercise problems with detailed solutions allow readers to test their knowledge. These problems can be used for practice or as templates to solve similar problems on actual hardware. Several of the problems require information from other
chapters and serve to further knit together various fundamental concepts. The solutions are available on the Artech House Web site at http://www.artechhouse.com/
static/reslib/thierauf/thierauf1.html.
The reader is assumed to have a basic background in electrical engineering,
including an elementary understanding of voltage, current, resistance, capacitance,
and inductance. No knowledge of signal integrity, electromagnetics, or transmission line theory is assumed. These subjects are introduced in the text as needed, often with graphs. The formulas are kept simple in those cases where mathematics is
used, and to avoid cluttering the text, mathematical derivations have intentionally
been omitted. References are provided for those wishing to understand the derivations, or the underlying physics. A catalog of particularly useful formulas has been
collected into one chapter at the end of the book for easy reference and to avoid
defocusing the earlier chapters. Those uninterested in using formulas can skip that
chapter, but those wishing to check the results from field and circuit simulations,
xv
xvi
Preface
or those wanting to use mathematics to validate their intuition, will find it a helpful
resource.
This book has greatly benefited thanks to the generous help from the following:
•
Simucad Design Automation for the SmartSpice license necessary to create a
number of the circuit simulations appearing in this book;
•
Simberian, Inc. for the 3D field solver SIMBEOR used to create and validate
all of the 3D circuit models used in the book;
•
Polar Instruments, Inc. for the Si9000 2D field solver that generated a number of the resistance, capacitance, and inductance results;
•
C&H Technical Sales and TTM Technologies for providing the trace photos
appearing in Chapters 5 and 8;
•
Tektronix, Inc. for providing the test equipment photograph appearing in
Chapter 13;
•
The AWR Corporation for the use of AWR® Microwave Office®;
•
Laraine Worby of the Framingham Public Library for obtaining many of the
obscure references cited;
•
The libraries of Wentworth Institute of Technology, Franklin W. Olin College
of Engineering, and Worcester Polytechnic Institute.
Once again, the enthusiastic assistance of the Artech House staff and the skillful review provided by the anonymous reviewer are most appreciated in helping to
turn a raw manuscript into a finished text.
I also thank my friend and colleague Jeff Cooper for calling to my attention the
ageless story of the first transatlantic cable mentioned in Chapter 1 and pointing
out its relevance to the problems faced by modern signal engineers and managers.
Most importantly, I am especially grateful to my wife Ann. Without her support, understanding, and encouragement, this book would not have been possible.
CHAPTER 1
Introduction to Signal Integrity
1.1 What Is Signal Integrity?
Signal integrity is the analysis, design, and validation of the interconnect necessary
for successful transmission of digital signals.
To do this, signal integrity (SI) engineering borrows ideas and practices from
many branches of electrical engineering. To those not having a background in analog or radio frequency electronics, this diversity can make SI seem mysterious,
sometimes inconsistent, and occasionally difficult to understand. Logic designers
accustomed to working with binary logic can find the apparently imprecise, analog
nature of SI particularly frustrating.
Although knowledge in RF and analog electronics is helpful, modern SI CAD
allows nearly anyone to create SI simulations and make predictions about circuit
operation. However, as Henry Petroski points out [1], in general CAD software
vendors do not always advertise or may not fully understand the limitations of
their products. When expertly used, CAD is powerful, but its misuse can have dire
consequences.
The SI engineer and technical manager working with pulse data transmission
can avoid pitfalls like this by understanding the fundamental ideas described in this
book.
1.2 The Importance of Signal Integrity: The First Transatlantic
Telegraph Cable
Managers and engineers not having a fundamental understanding of the technical
principles governing a project are more likely to make serious technical mistakes,
especially when pushing the state of the art. Perhaps the first time this became apparent with regards to a significant high-speed signaling project occurred on August
16, 1858. That was when Great Britain’s Queen Victoria sent a welcoming message
to U.S. President Buchanan over the first successfully laid transatlantic telegraph
cable [2–4]. This 3,800-km (about 2,400-mile) length of submarine cable connected
1
2
Introduction to Signal Integrity
Ireland with Newfoundland and could not be financially viable unless it operated at
speeds similar to dry land telegraph cables. Unfortunately, signal integrity problems
caused the Queen’s 99-word inaugural message to take nearly 17 hours for transmission instead of several minutes as had been expected [2].
The roots of this nineteenth-century disaster are both managerial and technical. The
underlying management fault was that the investors and managers did not possess a
rudimentary understanding of telegraphy. Because of this they hired a chief electrician lacking a thorough background in electrical engineering or telegraphy theory.
This decision was at the root of the cable’s technical flaws. Without this knowledge
the chief electrician was unable to understand the latest technical developments and
instead relied on his intuition to design the cable and instrument set. These intuitive
designs were appealing to management but inappropriate for the task.
Some outside experts (most notably Professor William Thompson, later to be
Lord Kelvin) had questioned the cable’s electrical design and the electrical equipment to be used with it. With his theoretical background, Thompson knew that
the electrical characteristics of a long submarine cable were significantly different
from the open air cable that had earlier been used in proof-of-concept experiments,
and he correctly believed that the submarine cable would distort high-speed Morse
code pulses. He preferred another cable design and suggested transmitting and receiving equipment with very different electrical characteristics than the equipment
designed by the chief electrician. Because of management’s weak technical background, they were unable to ask the kinds of probing, focused questions needed to
resolve the technical debate between Thompson and the chief electrician.
After several false starts, the intuitively designed but electrically flawed cable
was finally laid, and direct current from a battery in Ireland was detected by instruments in Newfoundland. This low-frequency test seemed to vindicate the design,
but it was soon discovered that high-speed pulses could not be properly received. In
fact, signaling was only possible at speeds that were far too slow to be profitable.
Management was unable to properly judge the merits of the various approaches to debug suggested by staff. Instead of using the sensitive transmitting and receiving instruments purposed by Thompson (which we now know would have
significantly improved the signaling speed), they authorized the chief electrician
to increase the voltage of the transmitted pulses. The misguided idea was that the
higher voltage would “push” the pulses down the cable with greater force, allowing signaling to occur at high speeds. In fact, this did not improve the signaling rate
because the fundamental problem was pulse distortion caused by the cable’s large
RC delay, not signal strength.
The large RC delay was the critical difference between the dry-land cables
used during the proof-of-concept experiments and the submarine cable. The dry-air
cables used air as insulation, but the submarine cable used a waterproof insulation
that greatly increased the capacitance. We will return to this effect in Section 1.7.
The large RC delay so distorted the pulses that only the slowest of them could
be properly received. Modern SI engineers would say that intersymbol interference
was created by the way in which the cable attenuated high frequencies (the highspeed Morse pulses) but allowed low frequencies (slow speed pulses, including
1.3 What Is a Pulse?
3
the battery’s DC) to pass. As we will see in later chapters, this same behavior also
occurs when signaling over modern circuit board traces.
During debug the sending voltage was incrementally increased to excessively
high values in an unfortunate attempt to improve signaling speed. This destroyed
the cable within a month of the Queen’s inaugural message. In fact, the cable never
sent any commercial telegrams and was a financial failure that set off investigations
of financial fraud and technical mismanagement and ruined several careers.
Modern SI engineers and those who manage them can avoid the kinds of systemic problems experienced during the 1858 cable project by learning the ways in
which pulses misbehave when traveling along long lines and learning techniques
for correcting these types of errors.
We begin by defining a pulse, then examine the time and frequency domains,
and describe two of the most important physical constants used in signal integrity
work. We conclude by briefly discussing the effects that many drivers simultaneously switching have on power supply noise.
1.3 What Is a Pulse?
A pulse is shown in Figure 1.1. The pulse width in time (W, usually nanoseconds
or picoseconds) is measured between the two points where the signal crosses half
of the signal amplitude (“the 50% points”), while the pulse period (T) is measured
at the pulse base. The rise time (tr) is defined as the time required for the pulse to
increase from 10% of its steady state value to 90% of that value. The fall time (not
shown in the figure) is measured between the same voltage levels on the falling edge.
The rise or fall time can be measured in many ways [5, 6], but in this book we will
stay with the commonly accepted “10/90” definition shown in the figure.
Under some conditions the pulse’s rising or falling edge may exceed and then
oscillate around the steady state value. Usually this is undesirable and later chapters describe the transmission line conditions that can cause the overshoot and
ringback conditions shown in the figure.
Overshoot
100%
90%
Ringback
50%
10%
W
tr
T
Figure 1.1 Pulse of width W, period T, and rise time tr overshoots and then rings back below the
steady state value.
4
1.4
Introduction to Signal Integrity
Time and Frequency Domains
Pulses described using time are said to be represented in the time domain. Oscilloscopes are common time-domain instruments, and the SPICE circuit simulator
displays its results in the time domain by default.
Most signal integrity work is performed in the time domain, but it is important
to understand the way in which the electrical characteristics of circuit board traces
change at different frequencies. This frequency domain approach is especially useful when studying how losses on transmission lines affect the shape (distortion) and
alignment of pulses (intersymbol interference).
A vector network analyzer (VNA), which displays the impedance of a component or transmission line at various frequencies, and the spectrum analyzer, a
device that simultaneously displays the amplitudes across a range of frequencies,
are examples of frequency-domain instruments.
1.4.1
Line Spectrums
From signal theory we know that a series of sine waves can be used to recreate a
recurring stream of pulses. In fact, a Fourier analysis can be used to find the amplitude and phase of each of the frequencies and determine how many are needed for
reconstructing the pulses to the desired level of precision [7, 8].
After doing this analysis, we find that there is a fundamental frequency that
has the highest amplitude and possible harmonic frequencies with lesser amplitudes
occurring at integer multiples of the fundamental frequency.
The fundamental frequency fo is found with (1.1) where T is the pulse period.
fo =
1
T
(1.1)
For instance, the fundamental frequency of a 10-ns stream of pulses is 100
MHz. The second harmonic (if present) is twice the fundamental frequency (200
MHz), the third is 300 MHz, and so on indefinitely. The amplitude of each harmonic depends on the duty cycle and rise time of the pulse, and some can have a
value of zero (that is, will not be present). None will have a value higher than the
fundamental. In fact, (1.1) shows the lowest frequency that will be present in a
pulse stream.
For example, a 10-ns square wave (a pulse with the same high and low times)
with zero rise time edges has a fundamental frequency of 100 MHz and only the
odd harmonics. This is represented in the frequency domain by a discrete line spectrum as shown in Figure 1.2. The amplitude of the fundamental has been set to 1,
and the harmonics have been scaled accordingly.
1.4.2 Recreating a Pulse with Sine Waves
The harmonics with the amplitudes shown in the spectrogram can be added together to recreate the pulse stream. In theory an infinite number of harmonics may
be necessary, but far fewer suffice in practice.
1.4 Time and Frequency Domains
5
1.0
Fundamental
frequency
Amplitude
0.8
0.6
3ed Harmonic
0.4
0.2
0.0
100
300
500
700
900
1,100
1,300
MHz
Figure 1.2 Line spectrum of a 10-ns square wave showing the fundamental frequency and the first
13 harmonics. For this pulse only the fundamental frequency and the odd harmonics have values
greater than zero.
Figure 1.3 shows how a 100-MHz stream of square waves can be reconstructed
by adding together a series of sine waves.
Panel A shows how the pulse looks when only the 100-MHz fundamental frequency is present. The period is 10 ns as expected from (1.1), but the signal is a sine
wave rather than a pulse.
Panel B shows the beginnings of a pulse when just the third harmonic is added to the fundamental. From the line spectrum we know the third harmonic is a
300MHz sine wave with an amplitude one-third of the fundamental.
Panel C shows the improvement when the fifth harmonic is added to the fundamental and the third harmonics. The waveform is clearly a pulse, but with large
ripple in the flat portion.
Panel D shows the further improvement when the seventh harmonic is added
to the previous results. Although the ripple is still present, it is higher in frequency
and been reduced in amplitude. This improvement process continues to infinity as
more odd harmonics are added.
1.4.3 Why Does the Frequency Domain Matter to Signal Integrity Engineers?
Instead of creating a pulse by adding together sine waves, we get a more powerful
insight by observing the effects when harmonics are removed. For example, if the
pulse shown in Panel D of the figure is launched though a filter that removes all
of the frequencies higher than the third harmonic, the results shown in Panel B are
obtained. This shows that by altering harmonics in the frequency domain, the pulse
is distorted in the time domain.
This observation is critical because lossy transmission lines act as lowpass filters and gradually remove higher frequencies. As we have just seen (and as the designer of the transatlantic telegraph cable belatedly discovered), this naturally leads
to pulse distortion. Losses are discussed in Chapter 9.
6
Introduction to Signal Integrity
0
5n
10n
15n
20n
0
5n
(a) fo only
0
5n
10n
(c) fo + h3 + h5
10n
15n
20n
15n
20n
(b) fo + h3
15n
20n
0
5n
10n
(d) fo + h3 + h5 + h7
Figure 1.3 Reconstructing a 10-ns pulse stream from the fundamental and the first three odd harmonics.
The progressive improvement in the pulse appearance is apparent as more harmonics are added.
A second relationship is shown in Figure 1.4. The solid circles represent the
spectra of the 100-MHz square wave when the rise time is zero. This is identical to
Figure 1.2 and is included for comparison with the hollow circles, which show the
spectra when the same pulse has a rise time of 1.5 ns. The fundamental frequencies
are seen to be the same, but the amplitude is lower for the 1.5-ns case, and some
even harmonics are now present (for instance, at 200 MHz and 400 MHz). We
further notice that for the 1.5-ns rise time case the amplitudes are essentially zero
for frequencies of roughly 600 MHz and higher. In contrast, when the rise time is
zero, the amplitude is still significant up to 1.3 GHz and beyond.
Apparently the zero rise time pulse requires more harmonics than an identical
pulse having a slower rise time. In the frequency domain this means the zero rise
time pulse requires more bandwidth than the 1.5-ns rise-time pulse for distortion
free transmission. Alternatively, we can say that less bandwidth is needed for distortion free transmission of a long rise time pulse than an identical, sharper pulse.
Pulse shaping (including rise time control) is an important aspect of high-speed
I/O design [9–12], but is outside the scope of this book. What is relevant to us is the
1.4 Time and Frequency Domains
7
1.0
Amplitude
0.8
0.6
0.4
0.2
0.0
100
300
500
700
900
1,100
1,300
MHz
Figure 1.4 Line spectrogram of a 100-MHz square wave pulse train when the rise time is 0 (solid
circles) and 1.5 ns (hollow circles).
effect that rise time has on transmission line behavior, and this fundamental topic
is discussed in subsequent chapters.
1.4.4 Upper Bandwidth
Equation (1.1) found the lowest frequency present in a pulse stream. The highest
frequency of interest is related to the signal rise time.
The rise time (tr) of a pulse exiting a lowpass filter (such as a circuit board
trace) is related to the filter’s 3-dB bandwidth (BW) by the relationship shown in
(1.2) [13]:
BW =
K
tr
(1.2)
The value of K depends on the shape of the rising or falling edge, and for
smooth, Gaussian type pulses when the overshoot is no more than 5%, it ranges
from 0.35 to 0.45 (depending on the waveforms characteristics) [14]. A value for K
of 0.35 is often used, but more conservative users choose 0.5 [6].
Although this equation shows how a lowpass filter will affect the rise time of
the output signal, SI engineers often use it to find the highest frequency of interest in a series of repeating pulses. For instance, to transmit a pulse with a 1.5-ns
rise time, the signal integrity engineer would say that the interconnect requires a
bandwidth no less than 233 MHz (using 0.35 for K). This is evident in Figure 1.4:
Although they have different values, all of the hollow circles for frequencies of 233
MHz and above have amplitudes below 0.6 (which is “3 dB down” from the value
of the fundamental). Working with decibels is discussed in Chapters 3, 9, and 16.
The spectrogram of a pulse having glitches or plateaus along its edges, or one
that overshoots or oscillates about the steady state value, will have more harmonics
8
Introduction to Signal Integrity
than the well-behaved pulses used here. Equation (1.2) should be used with caution
in these situations because those pulses require more bandwidth than the equation
predicts.
1.5 Power Integrity and Simultaneous Switching Noise
Power integrity is the analysis and design of power delivery systems. SI engineers
are sometimes expected to perform at least some power integrity tasks even though
these fields are separate and distinct. For instance, an SI engineer may perform
power supply decoupling analysis and power plane assignments within a stackup,
especially in small firms such as start-ups.
Although this book does not discuss power integrity in detail, simultaneous
switching noise is an important concept that the SI engineer must understand.
Those interested in learning more about power integrity are referred to [15, 16].
1.5.1 What Is Simultaneous Switching Noise?
Simultaneous switching noise (SSN) occurs when several output drivers simultaneously switch and draw current from a power supply [12, 17, 18]. The connections
for a typical power delivery network (PDN) are shown in Figure 1.5. The noise severity is determined by the PDN resistance, inductance, and decoupling capacitance.
Sometimes SSN is called deltaI noise to show that it is the change in current
that causes the power supply noise voltage. This term emphasizes that the inductance of the PDN is the root cause of SSN since the changing current through an
inductor determines the voltage drop across it. If the switching current cannot
be reduced (or the signal rise time lengthened), SSN is improved by lowering the
power supply inductance or adding decoupling capacitance (or both). As illustrated
R
L
Vdd
C
Vsupply
Output
drivers
Figure 1.5 Power delivery network resistance, inductance, and capacitance determines the value
of Vdd when I/O drivers switch.
1.6 What Is the Dielectric Constant and Loss Tangent?
9
in Figure 1.6, SSN affects the stability of the power source powering output drivers
and input receivers.
The power supply voltage remains reasonably steady when one driver switches,
but the increase in current occurring when many drivers switch causes a large voltage drop across the PDN inductance. This reduces the voltage at the transistors of
the I/O cell (Vdd in the figures). If the decoupling capacitance and resistance are
inadequate, the voltage will ring back up and exceed the steady state value when
one or more of the drivers stop drawing current. As diagrammed in Figure 1.6 with
CMOS I/O drivers this power supply variation changes the driver’s timing and often its drive strength. In fact, SSN is one of the major sources of timing jitter (see
[19–24]). The SI engineer must consider these factors when selecting I/O models
and setting up SI simulations (tasks described in Chapter 3).
1.6 What Is the Dielectric Constant and Loss Tangent?
Signal integrity engineers focus on two electrical parameters when comparing circuit board materials. The capacitance of circuit board traces is determined in part
by the value of the laminates relative dielectric constant, and the degree to which
the laminate contributes to signal loss is determined in part by the loss tangent.
The relative dielectric constant is abbreviated with the symbol εr. Often it is
simply called the dielectric constant and is sometimes abbreviated as Dk. The two
terms are synonymous, but circuit board shops and laminate manufactures generally use Dk rather than the more formal εr.
The relative dielectric constant of an insulator indicates by how much the capacitance increases when the dielectric (such as the laminate sheets between routing
layers), instead of air, is used to separate the plates of a capacitor. For example, a
(a)
Vdd
4V
1 Driver
3V
2V
5 Drivers
(b)
Driver output
4V
2V
1 Driver
5 Drivers
0V
Time
Figure 1.6 PDN inductance causes Vdd to drop below the 3.3-V steady state value when I/O drivers
switch. The drop is moderate when one driver switches, but is greater when 5 drivers switch simultaneously (a). The changing voltage affects the driver output timing [stylized in (b)].
10
Introduction to Signal Integrity
capacitor formed with a dielectric having εr = 2.6 (such as the Gutta-percha insulation used on the transatlantic submarine cable [25]) has 2.6 times the capacitance
of an identical capacitor formed in air (the insulation used in the transatlantic cable
proof-of-concept experiments). As Chapters 5 and 8 show, εr of laminates change
with frequency and they do not have a constant value.
The loss tangent (in this book abbreviated as LT, but also sometimes called the
dissipation factor, Df) shows the fraction of signal current lost in the dielectric and
so is not available to propagate the signal. This represents a frequency dependent
loss, and as we will see in Chapters 7 through 9, is an important factor in distorting
pulses when signaling at high frequency.
Loss tangent is often expressed as a decimal fraction, and dissipation factor
as a percent. For instance, a typical LT value for FR4 is 0.02, which can also be
presented as a Df of 2%.
Because it is a ratio of currents, the loss tangent has no units. The laminates
used in modern printed circuit boards have a low dielectric loss, but moisture uptake (described in Chapter 9) can cause dielectric losses to increase in high-humidity environments.
1.7
Main Points
•
Pulses may be examined in the time or frequency domains.
•
Pulses can be recreated by combining sine wave harmonics.
•
Passing a pulse through a lowpass filter such as a lossy transmission line removes high-frequency harmonics, distorting the pulse.
•
Pulses with small rise times require more harmonics (require a higher bandwidth) than a slow rise-time pulse.
•
Simultaneously switching output drivers cause power supply noise that can
distort the shape and add jitter to the output pulse.
•
The loss tangent indicates the amount of frequency dependent loss in the
dielectric that a signal will experience.
References
[1] Petroski, H., Remaking the World: Adventures in Engineering, New York: Alfred A. Knopp,
1997, pp. 70–71.
[2] Gordon, J. S., A Thread Across the Ocean, New York: Walker & Company, 2002.
[3] Standage, T., The Victorian Internet, New York: Walker and Company, 1998.
[4] Israel, P., From Machine Shop to Industrial Laboratory, Baltimore, MD: The Johns Hopkins
University Press, 1992.
[5] Ruston, H., and J. Bordogna, Electric Networks: Functions, Filters, Analysis, New York:
McGraw-Hill, 1966.
[6] Johnson, H., and M. Graham, High-Speed Digital Design, Upper Saddle River, NJ: PrenticeHall, 1993.
1.7
Main Points
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
11
Couch, L., Digital and Analog Communications Systems, 5th ed., Upper Saddle River, NJ:
Prentice-Hall, 1997
Poularikas, A. D., and S. Seely, Signals and Systems, Boston, MA: PWS Publishers, 1985.
Bateman, A., Digital Communications, Reading, MA: Addison-Wesley, 1999.
Bissell, C. C., and A. Chapman, Digital Signal Transmission, Cambridge, U.K.: Cambridge
University Press, 1996.
Kurzweil, J., An Introduction to Digital Communications, New York: John Wiley & Sons,
2000.
Dabral, S., et al., Basic ESD and I/O Design, New York: John Wiley & Sons, 1998.
Elmore, W. C., “The Transient Response of Damped Linear Networks with Particular Regard to Wide Band Amplifiers,” J. Appl. Physics, Vol. 19, January 1948, pp. 58–83.
Matick, R. E., Transmission Lines for Digital and Communications Networks, New York,
IEEE Press, 1969.
Novak, I., and J. R. Miller, Frequency-Domain Characterization of Power Distribution Networks, Norwood, MA: Artech House, 2007.
Bogatin, E., Signal and Power Integrity Simplified, Upper Saddle River, NJ: Prentice-Hall,
2009.
Hall, S. H., et al., High-Speed Digital System Design, New York: John Wiley & Sons, 2000.
Bakoglu, H. B., Circuits, Interconnections, and Packaging for VLSI, Reading, MA: Addison-Wesley, 1990.
Li, M. P., Jitter, Noise and Signal Integrity at High Speed, Upper Saddle River, NJ: PrenticeHall, 2008.
Neves, A. P., “Using Measure-Based Modeling to Analyze Backplane Deterministic Jitter,”
DesignCon 2004, Santa Clara, California, February 2–5, 2004.
Tektronix Corporation, “Understanding and Characterizing Timing Jitter,” Tektronix, Inc.
White Paper, 2003.
Pandit, V. S., et al., “SSO Noise, Eye Margin, and Jitter Characterization for I/O Power Integrity,” DesignCon 2009, Santa Clara, CA, February 2–5, 2009.
Miller, M., et al., “A Comparison of Methods for Estimating Total Jitter Concerning Precision, Accuracy and Robustness,” DesignCon 2004, Santa Clara, CA, January 29–February
1, 2007.
Li, M., “A New Jitter Classification Method Based on Statistical, Physical, and Spectroscopic Mechanisms,” DesignCon 2009, Santa Clara, CA, February 2–5, 2009.
Federal Telephone and Radio Corp., Reference Data for Radio Engineers, 3rd ed., New
York: Knickerbocker Printing Corp., 1949.
CHAPTER 2
The Signal Integrity Process
2.1
Introduction
Although modern signal integrity (SI) CAD is sufficiently sophisticated to predict
with great accuracy the shape and noise content of signal waveforms, this is not the
usual intent of SI analysis (advertisements from SI CAD companies notwithstanding). Instead, an SI analysis defines the boundaries of operation. By creating a group
of worst-case simulations, the SI engineer can confidently predict the absolutely
best and worst signal behavior and noise. No matter when or where it is manufactured, any given sample of the actual hardware is expected to fall within the bounds
defined by the worst-case analysis.
An experienced SI engineer, especially one who in the past has correlated SI
simulation results with laboratory measurements on hardware, will instinctively
know when a simulation contains too much margin or when the simulation results
are physically unreasonable. A less experienced SI engineer, particularly one with
no correlation experience, takes the simulation results more literally. These simulations are likely to show excessive margin or unrealistic signal characteristics.
There are special instances when the signaling environment is known in great
detail or where the electrical characteristics can be specified to within very small
tolerances. These cases often involve very high-performance signaling, or high-reliability systems signaling at more modest rates. In other instances electrical measurements of actual hardware can be used to fine-tune an existing SI model. In
these cases modern SI CAD will predict the actual signal wave shape and other
attributes with great precision.
2.2 Why Perform a Signal Integrity Analysis?
Perhaps the most common expected outcome of an SI analysis is to insure signal quality and validate system timing. Simulations are performed to quantify the
amount of timing margin and to determine that signals have adequate noise margin,
do not overshoot by too much, and have manageable crosstalk.
Signal integrity analysis can also determine if I/O drivers are subjected to stresses causing them to operate outside of their safe zone. Some stresses can permanently
13
14
The Signal Integrity Process
damage I/O circuits and shorten the components life. It is often more efficient to set
up and perform this type of analysis in simulation than by performing laboratory
measurements.
Other types of SI analysis include selecting between components to determine
the least expensive device for a particular application and to determine how a product can be cost-reduced. For instance, with SI simulations it is easy to decide with
great confidence if an expensive close tolerance termination resistor or expensive
type of decoupling capacitor is required in a particular design. It is also straightforward to observe the benefits of various circuit board stackups and materials or, in
an ASIC, to select between various I/O drive strengths.
Because well-managed signals generally have lower harmonic content than
poorly terminated signals, a common axiom is that good SI yields good electromagnetic interference (EMI) test results. Although good SI is not a guarantee that a
product will flawlessly pass EMI testing, experience shows that favorable EMI test
results are usually significantly less likely when SI problems are present.
In some instances SI simulations are not performed until after the product has
arrived in the laboratory, generally as part of a tardy effort to understand unpredictable system behavior. This can be done on a prototype set aside for debug
purposes or on a device already in production that has been identified by manufacturing as having SI problems. Performing an SI analysis this late in the product
design cycle is inefficient and expensive, but it can identify the least costly method
of improving signal quality.
2.3 What Is a Typical Signal Integrity Workflow?
An idealized workflow performed by one or more signal integrity engineers is illustrated in Figure 2.1. Although eight distinct phases are shown, adjacent phases
are sometimes merged, and not all projects need to carry out all of these tasks. Additionally, the interaction between phases is not shown [for instance, the continual
feedback between debug and design verification testing (DVT)].
During the product’s architectural phase, a senior, experienced SI engineer
identifies the limits of technology (such as circuit board size or thickness or ASIC
signaling speeds) and broadly defines the circuitry requiring detailed analysis. New
CAD tools are also evaluated and purchased at this time, including 2D and 3D field
solvers [1]. Decisions are made as to the type of models (transistor level or IBIS
circuit models, lossy transmission line, or S-parameter models, for example) that
will be used. A rough plan describing the power distribution network is developed.
This phase is often merged with the prelayout phase, especially if the design
firm is small (such as a start-up) or has experience with similar technology (and
so already has the critical circuit models), or when the product is not aggressively
pushing new technology.
The prelayout phase is the time when SI models are obtained and validated
and CAD tools are set up. Compatibility between models (verifying that models
from different vendors operate correctly when used in the same simulation) is also
tested [2].
The validation process individually tests the circuit model of each component by
comparing simulation results under simple conditions to laboratory measurements
2.3 What Is a Typical Signal Integrity Workflow?
Architecture
Prelayout
15
• Feasibility studies
• Broad bound limits
• Identify critical nets
• Define layout constraints
Layout
• Confirm placements
• Monitor constraint compliance
Postlayout
• Confirm constraint compliance
• Validate critical nets
Release to manufacturing • Fabricate and assemble prototypes
DVT
• Perform margin testing
• Identify bugs
Debug
• Fix bugs
• Update documentation
Release to production
• Ship product to customers
Figure 2.1 Idealized signal integrity workflow showing eight distinct phases. Interaction between
phases is not shown.
or to results known to be correct. This time-consuming and often overlooked (and
unscheduled) task is vital for proper correlation between SI simulation results and
the performance of actual hardware.
Critical nets (and classes of nets) are identified by simulation, and layout constraints (including assumed parts placement) are developed [3]. The need for and
the style of termination are determined, and critical layout rules (such as trace
widths, spacing, and impedance values) are recorded in an SI layout rules document that will be given to the layout designer. If need be, this is when simulations
with advanced laminate systems are carried out to improve signal quality or routing density and the laminate is selected. The circuit board fabrication shop develops a stackup showing the thickness of each layer, the placements of planes, and
the trace widths necessary to obtain the specified impedance. The analysis of the
power delivery system is begun (sometimes by power delivery specialists or EMI
engineers, but often by signal integrity engineers).
During the layout phase, the layout designer creates the artwork necessary to
create the circuit board. During this time the SI engineer periodically reviews the
circuit board artwork and often continues with prelayout simulations. This is also
a time to determine if the manufacturers have updated the circuit models for critical integrated circuits and connectors.
The SI engineer often finds the postlayout phase to be the most hectic of the entire design cycle. This is when the finished artwork is examined for compliance with
the SI rules, and when constraints such as spacing between traces, trace widths, and
lengths and termination placement and timing are checked. Trade-offs are made to
accommodate physical constraints, and simulations are rerun to confirm proper
16
The Signal Integrity Process
operation. Some CAD tools support postlayout analysis, so if properly set up during prelayout, the completed layout database can quickly be automatically analyzed for compliance and timing. Although this provides an enormous time savings,
especially on large multilayer boards, these tools are no substitute for a detailed
visual inspection of the artwork (and the SI engineer should always insist on having
enough time in the schedule to do this). The layout designer changes those things
identified by the CAD and visual inspections, and the SI engineer reexamines the
artwork. This process continues until the board fully meets the SI and timing requirements for the product.
The layout designer formally releases the board to manufacturing once the
artwork has been corrected. The SI engineer uses this time to finish up any final
simulations and to finalize preparations for design verification testing (DVT).
Generally, DVT is carried out by the engineer responsible for the circuit board,
and not the SI engineer. Margin testing where voltage and temperatures are set to
extreme values is preformed during this phase, and bugs are uncovered. Test patterns are used to exercise those highest risk nets identified during the prelayout and
layout phases. Often this phase is not distinct from debug, but some large firms
segregate these tasks and have different specialists for each. As described here, the
DVT process includes industry compliance testing, where those signals that must
strictly adhere to industry specifications are checked. In some cases, a third-party
testing laboratory is used to certify compliance.
Many SI engineers are not directly involved with laboratory debug. Instead,
they provide debug support by analyzing measurements taken by others and in
identifying critical test patterns and nets. This is especially likely if the SI engineer
is a CAD specialist and not one comfortable performing laboratory measurements.
However, the most valuable SI engineer is skilled in both CAD and measurement
techniques. These engineers have practical experience that is invaluable when interpreting simulation and measurement results.
The system is formally released to production for volume manufacturing once
the product has successfully passed through DVT and debug. Generally the SI engineer is not involved in solving production problems, but they may be required to
help resolve timing or signal quality problems that arise during volume production
and are often required to help with component selection or evaluation when critical parts become unavailable. This is especially so if the SI engineer has laboratory
measurement experience, but SI CAD specialists are also enlisted. If properly documented and archived, the prelayout and postlayout design analysis (particularly
test patterns and environmental setup assumptions) can be immensely helpful in
solving urgent, unexpected production SI problems.
2.4 Signal Integrity Worst-Case Analysis
Often the SI simulations performed during the architectural phase are done under
nominal conditions. Circuit models for typical silicon are used, and traces, power
supply voltages, and termination resistor values are all assumed to be exactly as
specified. As the design progresses, the analysis progressively takes into account
process variations in the components, power supplies, and circuit board traces.
2.4 Signal Integrity Worst-Case Analysis
17
Circuit board vendors do not specify the manufacturing tolerances of the trace
widths or thicknesses, but on those traces specified as controlled impedance they
do indicate the impedance tolerance. This range is used by the SI engineer in simulations to determine the optimum value of termination resistors and to find the
highest ringback and overshoot voltages.
2.4.1 Silicon Worst-Case I/O Models
Generally silicon vendors bound the range over which their device will operate by
providing nominal, fast (sometimes called best-case), and slow (sometimes called
worst-case) I/O circuit models. Because I/O driver performance is strongly dependent on the die temperature (the junction temperature) and the actual power supply
voltage, these external factors must be appropriately set along with the proper I/O
case for a specific type of analysis. For example, CMOS I/O launches the slowest
signal rise time when the lowest power supply voltage and highest temperature is
used with the slowest case I/O model. To determine the highest current draw the
fast I/O model is used with a low die temperature, highest power supply voltage and
lowest transmission line impedance.
The temperature parameter used in the I/O models represents the junction temperature, not the temperature of the ambient air. Sometimes this distinction can be
overlooked when the integrated circuit is a low power device, but it becomes critical as the device power dissipation increases. For instance, a commercial product
may operate in an ambient air temperature of 85°C or less. The junction temperature will not be much higher than this if the integrated circuit is low power device,
but (depending on the packaging) under the same circumstances the junction temperature of a higher power device may exceed 125°C. The junction temperature is
estimated during prelayout and this value is used in the evolving SI model.
Although the junction temperature can be set in the simulation, not all CAD
tools have the ability to individually set the junction temperatures of various integrated circuits simultaneously coexisting in a simulation. When this is the case,
a net containing a device operating at a high junction temperature connected to a
device with a lower junction temperature will be simulated at one temperature (often the highest one). The result is that Idd noise (noise on the power supply) and the
rise time of the signals launched by the I/O drivers will be underestimated for the
device operating at the lower temperature. This can mask signal integrity problems
such as the magnitude of overshoots and reflections.
2.4.2 What Combination of Environmental Effects Are Worst Case?
The precise combination of environmental conditions and circuit models that produce a particular worst-case condition depends on the details of a design, and sometimes the combination can be counterintuitive.
Rather than manually running through the various permutations, many SI
CAD tools can be set up to vary the parameters automatically and summarize the
results. This can be very helpful when designing large and complex circuit boards,
but adequate time should be included in the schedule to interpret and react to the
results.
18
The Signal Integrity Process
For instance, particularly surprising results should be examined (and possibility manually resimulated) to determine if they are valid or if they are the artifact of a
defective circuit model. The validity of the stimulus and environmental conditions
should also be checked to verify that the conditions presented in the simulation is
physically possible in the actual hardware.
Examples of some common worst case configurations are shown in Table 2.1,
where L and H represent the lowest and highest values specified by the manufacturer, F is the fast silicon circuit model, and O and E are the odd and even switching
modes (described in Chapter 9). The device current is shown as Idd, and the worst
signal overshoot is indicated in the highest OS column.
In general, setting the environmental variable to opposite values than shown in
the table will produce the opposite effect. For instance, to find the lowest Idd, use
the highest temperature, the lowest Vdd, the slow silicon process file, and so on.
2.4.3 Using SI Analysis Results During Debug
The results from the SI analysis are valuable when setting up laboratory conditions
for debug. For instance, noise between the power and ground planes will generally
be the highest when Idd is high and the signal rise time is short. Table 2.1 shows
that to make this measurement, integrated circuits with the fastest rise time I/Os are
placed on the circuit board, since that implies they are from the fast process corner.
The power supply voltage is raised to the highest permissible value, and the temperature is lowered to the lowest possible ambient value. If the testing is done on a
single circuit board or small assembly the temperature can be controlled by placing
it in a thermal chamber.
The signaling patterns identified during the SI analysis are useful during debug for exercising prototype hardware. By working closely with the diagnostic
engineer (a specialist responsible for creating the software and firmware necessary
to troubleshoot and diagnose faults in the hardware), the signal integrity engineer
can have the most critical SI test patterns included in with the test programs. In
this way tests that were run in simulation can be duplicated in prototype hardware
and the results compared. Besides validating the design, this correlation exercise
provides an excellent opportunity for the SI engineer to understand the limits of
simulation and to refine his or her intuition.
Table 2.1
Variable
Example Worst-Case Settings
Highest
Fastest
Highest
Highest Idd Coupling Rise Time OS
Temperature
L
L
L
L
Vdd
H
H
H
H
Silicon
F
F
F
F
TLINE Zo
L
H
H
H
Rs
L
L
L
L
Rt
L
H
H
H
Pattern
O
E
E
E
2.5
Main Points
19
2.4.4 Electrical Overstresses
Electrical stresses on an I/O pin can be caused by electrostatic discharge (ESD)
events and by signal pulses high enough in amplitude to damage the junctions and
gate oxide of the transistors in the I/O cell.
ESD testing is not included in with the normal SI simulations, but signal overvoltages must be analyzed carefully.
Time-dependent dielectric breakdown (TDDB) [4, 5] and hot carrier effect
(HCE) [5–8] stresses are present in the transistors forming the receiver and driver
portions of an I/O cell in response to high voltages at the I/O pin. Over time these
failure mechanisms can cause a gradual degradation of the transistors (changing
their behavior), or if the overstress is large enough, it can cause the transistors
to fail catastrophically. High voltage is a relative term and the actual maximum
voltages may only be a few volts higher than the maximum allowed power supply voltage. The specific circuit design of the I/O cell and the details of the process
technology determine the highest allowed voltage and the amount of time that a
particular voltage can be tolerated [4].
The signal integrity simulations should always be carefully examined to determine the highest voltage present at the I/O pin, and this value must never be
allowed to exceed the manufacturer’s worst-case specification. Both ends of the net
should be examined.
The highest value is usually created by large reflections generated when the
transmission line is not properly terminated, but they can also be created by crosstalk on high-impedance nets. Often the worst case occurs when the power supply
is at a maximum value, the temperature is low, and the series termination has its
lowest value since these conditions encourage the highest reflections. These failure
mechanisms are usually not a practical concern in high-speed serial signaling because there the signaling voltages are low and the lines are well terminated.
Often the manufacturer’s data sheet will show that an I/O driver is more able to
tolerate a high-going voltage than one going below ground, so special care should
be taken to design simulations to test the case where signals significantly drop below ground. Once identified, the simulation case creating the worst-case voltage
should be noted and that same case duplicated in the laboratory during debug or
DVT to verify that the maximum limits are not exceeded.
2.5
Main Points
•
SI simulations bound the range of operation and show how the system will
behave under all legitimate environmental conditions and manufacturing
variations.
•
There are eight distinct places in the design cycle where SI analysis can be
performed, but not all are performed on a specific product or by an individual engineer.
•
Good EMI test results begin with good SI.
•
Performing SI analysis late, after the product has been released to production, is costly and time-consuming.
20
The Signal Integrity Process
•
Patterns and environmental configurations identified during simulation can
be used to test actual hardware during DVT and debug.
•
SI analysis identifies design vulnerabilities in the design and shows where the
debug activities should be targeted.
References
[1]
Huang, C. C., “Signal Integrity Modeling and Simulation Tools,” Proceedings of DesignCon, 2004.
[2] Beal, W., “Use the Right Models for the Simulation of Multi-Gigabit Channels,” Proceedings of DesignCon, 2004.
[3] Ramakrishnan, P., “CDNLive! Paper—Signal Integrity (SI) for Dual Data Rate (DDR) Interface,” Cadence Designer Network, Silicon Valley, 2007.
[4] Chandrakasan, A., et al., Design of High-Performance Microprocessor Circuits, New
York: IEEE Press, 2001.
[5] Pompl, T., and M. Kerber, “Failure Distributions of Successive Dielectric Breakdown
Events,” IEEE Transactions on Device and Materials Reliability, Vol. 4, No. 2, June 2004.
[6] Leblebici, Y., and S. M. Kang, Hot-Carrier Reliability of MOS VLSI Circuits, Boston, MA:
Kluwer Academic Publishers, 1993.
[7] Dillinger, T. E., VLSI Engineering, Upper Saddle River, NJ: Prentice-Hall, 1988.
[8] Groeseneken, G, “Hot Carrier Degradation and ESD in Submicrometer CMOS Technologies: How Do They Interact?” IEEE Transactions on Device and Materials Reliability, Vol.
1, No. 1, March 2001.
CHAPTER 3
Signal Integrity CAD and Models
3.1
Introduction
The signal integrity (SI) engineer uses a variety of CAD tools and model types
when performing an analysis. This chapter introduces the two most common types
of models for representing integrated circuit input and output (I/O) drivers and
discusses some of the circuit models used to predict trace and via behavior. Finally, the field solvers necessary to create circuit models from physical structures
are described.
3.2
I/O Models
The models of the I/O circuits used in SI simulations are either transistor level circuit models or behavioral type models. The most prevalent circuit models are based
on the SPICE circuit simulator, while IBIS is the most common behavioral model.
3.2.1 What Are Transistor Level Models?
Because they come directly from the integrated circuit net list, transistor level models are the most accurate representation of the I/O circuitry. However, the accuracy
of the system level simulation also depends on the accuracy of the models used to
simulate the transistors (the worst-case I/O models discussed in Chapter 2), and the
quality of the models for the interconnect within the micropackage that houses the
integrated circuit.
To find the current and voltage at the I/O cell pad, the circuit simulator must
solve the equations for each transistor, capacitor, and resistor within the cell at
each time point in the simulation. While this can give highly accurate results, these
calculations are time-consuming and make transistor level simulations run much
more slowly than behavioral model simulations. In some instances the simulator
may have difficulty properly calculating the circuit node voltages and currents.
The simulation will prematurely terminate if the errors associated with these calculations become large enough. Circuit designers use various optional simulator
settings to minimize or avoid these kinds of problems, and occasionally different
21
22
Signal Integrity CAD and Models
I/O models will require that the same simulator option be set to separate values.
This makes it impossible to run a signal integrity simulation that simultaneously
includes both integrated circuits, since to do so would require the same parameter
to simultaneously have different values. Conflicts of this type are more likely to occur when the integrated circuits come from different vendors.
Because a transistor level I/O model is a net list that includes all the connections and values for all the components in the I/O circuit, it is possible to reverseengineer the model and discover exactly how the circuit works. To prevent this,
the I/O circuitry (and sometimes also the micropackage) can be encrypted by the
manufacturer. This hides the internal wiring and the transistor parameters, making
it impossible for anyone to determine the transistor process information or how
the circuitry is wired. The encrypted models are not compatible with all simulators,
so the SI engineer must confirm that an encrypted model will properly run on the
circuit simulator of their choice (and with the other models in the system).
Although encryption solves the manufacturers’ intellectual property dilemma,
simulations with encrypted models can be difficult to debug because the SI engineer
cannot examine or initialize internal nodes when the simulation goes awry.
3.2.2 What Are IBIS Models?
I/O models based on the ANSI/EIA-656-B standard (called IBIS, for I/O Buffer
Information Specification [1]) are the most popular behavioral type models. IBIS
models present the I/O terminal current and voltage values in look-up tables. The
simulator uses this information to determine the response of input and output circuits to different loads and voltages present at the I/O pin. These models include
estimates for the package parasitics, so a separate micropackage model is not necessary. The capacitance of the I/O circuitry is also included.
Because the simulator does not perform the complex calculations necessary for
transistor level models, simulations using IBIS models run significantly faster than
transistor level simulations.
Since no proprietary information is present, IBIS models are not encrypted, and
IBIS models can be run on simulators from many different vendors. However, not
all IBIS models fully comply with the industry specification, and some simulators
may not operate with models from all vendors.
A disadvantage of earlier IBIS models was that current in the power and ground
pins is not modeled, thus preventing these models from determining power supply
noise. A second problem of importance in high-speed signaling is that complex I/O
behavior such as precompensation and equalization (briefly described in Chapters
13 and 16, but more fully described in [2–8]) cannot be properly modeled. However, models conforming to the revision 5 IBIS specification (IBIS-AMI, for IBISAlgorithmic Modeling Interface [1]) remove these deficiencies.
These advanced features are optional and may not be included in all version
5 models. For instance, a model complying with the version 5 specification may
deliberately choose not include the power and ground currents but to include the
precompensation behavior. Each IBIS model should be checked to insure it has the
proper components for a given analysis.
3.3 Modeling Transmission Lines
23
3.3 Modeling Transmission Lines
Short transmission lines are sometimes modeled as a string of resistors, capacitors,
and inductors. Models of this type are occasionally provided by vendors for small
structures such as connectors, and signal integrity engineers sometimes create these
models for vias.
The resistors represent conductor and dielectric signal loss, but because the resistors are fixed values they do not change with frequency. As described in Chapters
6–8, this is undesirable and makes the model most accurate for single frequency
sine waves rather than pulses (which contain many frequencies).
Virtually all modern versions of SPICE include lossless and single-frequency
lossy transmission line models, and some SPICE implementations have additional
models. These are much more computationally efficient and convenient than concatenating strings of RLC circuits. The most common transmission line models are
listed in Table 3.1.
The lossless model (the SPICE T element) is a perfect transmission line comprised of only an inductance (L) and capacitance (C) and so its only characteristics
are delay and impedance. It can be accurately used with high-frequency pulses
when the losses are low (this means that the transmission line is electrically short).
In most implementations of SPICE, the T element models only one signal conductor,
so coupling between traces (crosstalk) and differential pairs cannot be simulated.
The lossy line (usually referred to as an LTRA line and often referenced as the
O element) is defined by specifying values for the capacitance, inductance, series
resistance (representing the loss in the conductor), and in some implementations a
shunt conductance (representing the dielectric losses). The resistance (R) and conductance (G) are fixed values that do not change with frequency, making this line
most accurate at the one frequency for which R and G were specified. As such, it is
best used with single frequency sine waves and will not be accurate when used with
digital pulses. Like the T element, in most implementations the O element models
one signal and cannot be used to simulate crosstalk or differential pairs.
At least two vendors [9, 10] provide a multiconductor lossy transmission line
element (often called the W element) where the circuit values automatically change
with frequency. The parameters in this model can be specified in many ways, but
the most common usage at low to moderately high frequencies is for the SI engineer
to provide values for L, C, R, and G. This model differs from the O element in that
the simulator computes new values for the conductor and dielectric losses (R and
G) and optionally for L and C at the various frequencies present in the signal. This
allows the model to properly account for losses in pulse streams where the signal
rise time and pulse width change (as can be caused by signal reflections). This
Table 3.1 Common SPICE Transmission Line Elements
SPICE
Element Type
Notes
T
LC Lossless
O
RLGC Lossy Losses correct at only one frequency
Only delay and impedance; no losses
W
RLGC Lossy R, L, G, C can be made to vary with frequency
24
Signal Integrity CAD and Models
model can simulate a single conductor or a multiconductor bus and so can be used
to model crosstalk or differential pairs.
The Exercises for Chapter 7 show how to compute R and G for use in O and
W element circuit models.
3.4
S-Parameters
Interconnect (including traces, vias, or connector pins) can be modeled in the frequency domain with scattering parameters (S-parameters). Although they originated in RF engineering, S-parameters are now commonly used in high-speed digital
signaling work, especially in serial signaling applications.
Because of their importance to advanced SI applications, an overview is presented here, but since S-parameters are not used in this book, we will not go into
detail. Additional introductory information on S-parameters is presented in [11–
13]. More theoretical explanations are presented in [14–17].
3.4.1
What Are S-Parameters?
S-parameters show the degree to which sine waves are reflected (scattered) from a
signal source of known impedance (the reference impedance) by each pair of terminals (ports), and the amount of the signal appearing between ports. This means that
S-parameter models show loss, impedance mismatch (reflections), and the coupling
between ports (crosstalk between traces).
A trace over a ground plane has two pairs of terminals and so is an example of
a simple two-port network, as shown in Figure 3.1.
Subscripts are used to indicate the terminal pair being measured (identified by
the subscript m) and the terminal pair injecting the signal (the subscript i). This is
shown in (3.1).
Smi
(3.1)
For instance, measuring the signal at port 2 and injecting a signal into port 1
gives S21 (the insertion loss, since it shows how much the signal is attenuated between ports 1 and 2).
Simultaneously injecting a wave and observing the reflection at port 1 yields
S11 (the reflection loss because it shows how much of the signal is reflected back
to the generator). Interconnects such as transmission lines and vias have the same
behavior in both directions, so for traces S21 = S12 and S11 = S22.
In S-parameter models the S-parameter coefficients are tabulated for a range
of frequencies and are represented by two numbers, one for the amplitude and the
other for the phase. A sample S-Parameter listing of just the amplitude for a 4-mil
(0.11-mm)-wide half-ounce 50 stripline on FR4 is shown in Table 3.2.
3.4.2 What Is Insertion Loss?
For a lossless transmission line, the insertion loss (S21) and its twin (S12) are 0, since
that indicates that no signal energy is lost when the transmission line is inserted
3.4
S-Parameters
25
Port 2
RL
Trace
Ground
plane
Laminate
Port 1
Vs
RR
S 21
RR
Vs
Port 1
Port 2
RL
S 11
Figure 3.1 Single trace over a ground plane two-port network. For S11 port 1 is simultaneously
driven and measured at port 1. S21 is measured at port 2 when port 1 is driven. In both instances
port 2 is terminated and driven with a reference impedance (RR).
Table 3.2 S-Parameter Coefficients
in Decibels for a 15-cm (6-Inch)-Long
50Ω Stripline as Calculated by Linpar
[18]
GHz
S11
S21
S12
S22
1.0
−56
−1.3
−1.3
−56
1.5
−51
−2.2
−2.2
−51
2.0
−46
−3.2
−3.2
−46
2.5
−42
−4.7
−4.7
−42
between the transmitter and the load. From the table we see that the insertion loss
for the 4-mil-wide 50Ω trace is actually −1.3 dB at 1 GHz and increases to −4.7 dB
at 2.5 GHz. The use of decibels is described in Chapters 8 and 17, where we will see
that a negative value indicates a loss and larger negative values represent larger loss.
26
Signal Integrity CAD and Models
Here we note that a loss of 1.3 dB means that the output is only about 86% of the
input value. This means that about 14% of the signal has been attenuated. About
41% of the signal is lost at 4.7 dB.
Chapter 1 showed that the 3-dB bandwidth is a convenient way to estimate
the highest frequency of interest in a pulse stream. From Table 3.2 we see that frequencies of 2 GHz and higher attenuate the signal by more than 3 dB. Therefore,
the bandwidth for this trace is just under 2 GHz, and signals (including the signal
harmonics) with frequencies higher than this will be severely attenuated.
3.4.3 What Is Reflection Loss?
The reflection loss (S22 and S11) indirectly shows by how much the transmission line
impedance differs from the reference impedance. Chapter 11 shows that a perfect
match results in no reflections, which would make the reflection loss zero. Since on
a decibel scale a large negative number represents a small quantity, a large negative
reflection loss value in decibels shows that the reference impedance and transmission line impedances are very similar. For instance, a reflection loss of −26 dB represents a 5% impedance mismatch, which is acceptable in most low and moderate
speed applications. From Table 3.2 we see that at 2 GHz (the 3-dB bandwidth of
this trace) the reflection loss is much smaller than this (−46 dB). That represents
about a 0.5% mismatch in the impedance.
3.4.4 What Are Touchstone Files?
Those versions of SPICE capable of running S-parameter models nearly universally
use the Touchstone File format [19] to present the S-parameter data, although other
formats also may be supported. Touchstone format model files can come from test
equipment (measurements) or field solvers (simulations).
Often the data is presented in Touchstone files using a magnitude-angle format
with the magnitude in decibels (dB) and the phase angle measured in degrees, but
the specification also defines other formats.
Usually S-parameters are specified at 50Ω, but any reference impedance may be
used and its value recorded in the file.
3.5
Field Solvers
Electromagnetic field solvers analyze a physical structure such as a circuit board
trace or the complex geometry of a connector and produce an electrical circuit
model for use in signal integrity simulations.
The most common (and often the easiest to use) field solvers are those that
analyze two-dimensional structures (2D solvers). Three-dimensional structures are
analyzed with 3D simulators.
3.5.1 What Are 2D Field Solvers?
The underlying assumption behind 2D solvers is that the structures shape is uniform
and so can be properly described with only two dimensions. A group of straight
3.5
Field Solvers
27
circuit board traces that have a constant width, spacing, height above the ground
plane, and thickness is a good example of the type of structure appropriate for 2D
analysis.
The field solver calculates the resistance, capacitance, inductance, and conductance of the 2D structure and tabulates them across a user-defined frequency range.
The frequency range must be carefully selected because, as we saw in Chapter 1,
digital pulses contain high-frequency harmonics and (as we will see in Chapters 6
through 8) losses increase with frequency.
Since the shape is uniform, the electrical parameters are specified per unit
length. The value for a specific trace is found by multiplying the field solver results
by the actual length of the structure. For example, if the capacitance of a trace is
reported as 140 pF/m, the capacitance for a 1-cm-long trace is 1.4 pF.
The best solvers allow the user to choose the output formant from a variety
of circuit model types, so a library of circuit models can be created that are usable
by a number of different circuit simulators. At a minimum the field solver that
you choose should produce both an RLGC output file and an S-parameter model
(preferably in Touchstone format). The RLGC data of single traces can be validated by comparing it with the results from the approximate formulas presented in
Chapter 17. This is a good way to verify that the model has been setup correctly.
As described in Chapter 7’s Problems, the RLGC information can be used to create
a W line model in those cases where the field solver does not directly create one.
Alternatively, it can be used with the SPICE LTRA model (noting the limitations
discussed in Section 3.3).
3.5.2 What Are 3D Field Solvers?
Circuit models for physical structures that are not uniform in shape (such as vias or
traces that vary in width or spacing or that bend around objects) are created with a
3D field solver. Some of these solvers require knowledge of radio frequency terms
and concepts to produce valid models, but other simulators are specifically designed
to be operated by signal integrity engineers having little RF knowledge.
Because the structure is simulated in three dimensions, setting up the problem
space is usually more time-consuming and error-prone than with 2D simulators.
Before purchasing a 3D simulator, check that entering the circuit board stackup and defining the material properties are intuitive. Furthermore, it should be
straightforward to verify that the values have been assigned properly and assigned
to the correct layer in the stackup. Surface roughness (as shown in Chapter 8, an
important factor when calculating conductor loss) should be easy to apply to traces
and planes. When working with high-speed signals, the ability of the simulator to
include the way in which frequency changes the dielectrics electrical properties is
also very important.
As part of the selection process, the drawing interface should be thoroughly
tested to determine the ease in creating and editing complex 3D structures. Although some 3D simulators can directly import the 3D mechanical drawings created by CAD drafting software, or the stackup from circuit board artwork tools,
it is routinely necessary for the SI engineer to “clean up” the drawing and to crop
regions once it has been imported.
28
Signal Integrity CAD and Models
Some 3D solvers have wizards or template files that make it easy to setup commonly found problems such as vias or differential pair traces (or differential vias).
This valuable feature can save significant time and is especially important if the SI
engineer is not using the simulator on a daily basis.
Generally S-parameters are the default output model type, but RLGC files or
W line models can also sometimes be created. The electrical parameters calculated
are of the complete geometry. For instance, a structure might have a capacitance of
3 pF, but it would not have a value of 3 pF/inch.
3.6
Main Points
•
I/O models are either transistor based or behavioral.
•
Transistor level models are most accurate but run significantly slower than
IBIS models.
•
Transistor level models may be encrypted to protect intellectual property.
This can make the SI debug more difficult.
•
IBIS models are the most common behavioral type model and run significantly faster than transistor level models.
•
Field solvers calculate the resistance, capacitance, inductance, and conductance of physical structures.
•
Two-dimensional field solvers are used to create circuit models of regular
structures such as uniform traces.
•
Three-dimensional field solvers are used to create circuit models such as vias
and irregular structures.
•
Some 3D simulators are difficult to set up and to properly run and should be
thoroughly tested to verify that the operation is intuitive.
•
The output from 2D and 3D field solvers can be tabular data, circuit models
of various forms, or S-parameters.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
IBIS version 5.0, ratified August 2008 by the IBIS Open Forum, http://www.eigroup.org/
ibis/specs.htm.
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Proakis, J. G., Digital Communications, New York: McGraw-Hill, 2000.
Bateman, A., Digital Communications, Reading, MA: Addison-Wesley, 1998.
Bissell, C. C., et al., Digital Signal Transmission, Cambridge: Cambridge University Press,
1996.
Horowitz, M., et al., “High-Speed Electrical Signaling: Overview and Limitations,” IEEE
Micro., Vol. 18, No. 1, January/February 1998.
Dally, W., et al., “Transmitter Equalization for 4-Gbps Signaling,” IEEE Micro, Vol. 17,
No. 1, January/February 1997.
3.6
Main Points
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
29
Dally, W., et al., Digital Systems Engineering, New York: Cambridge University Press,
1998.
SmartSpice, Simucad Design Automation, Santa Clara, CA.
HSPICE, Synopsys, Inc., Mountain View, CA.
Quint, D., and K. Bois, “The Digital Designer’s Complete Lossy Transmission Line Model,”
Proceedings of the 52nd Electronic Components and Technology Conference, May 28–31,
2002, pp. 1073–1079.
Hewlett Packard Test & Measurement Application Note 95-1, “S-Parameter Techniques
for Faster, More Accurate Network Design,” November 1996.
Bogatin, E., “How Do You Spell S Parameters?” Printed Circuit Design Magazine, February 2003.
Hall, S. H., et al., High-Speed Digital System Design, New York: John Wiley & Sons, 2000.
Young, B., Digital Signal Integrity, Upper Saddle River, NJ: Prentice-Hall, 2001.
Agilent AN 154, “S-Parameter Design,” Application Note, Agilent Technologies, 2000.
De Geest, J., et al., “Electrical Modeling of High Speed Interconnection Links,” Proceedings of the 51st Electronic Components and Technology Conference, May 29–June 1,
2001, pp. 1245–1249.
Djordjevic, A. R., et al., LINPAR for Windows, Norwood, MA: Artech House, 1999.
Touchstone Specification, Version 2.0, ratified April 24 2009 by the IBIS Forum, http://
www.eigroup.org/ibis/specs.htm.
CHAPTER 4
Printed Test and Evaluation Boards
4.1
Introduction
Usually debug is performed in the laboratory on prototype versions of the actual
product board. This allows hardware, software, and firmware developers to perform debug on hardware that is essentially identical to what will become the final
product.
However, this approach is not always the most efficient, especially when working with new signaling or integrated circuit technologies and new devices (such as
connectors or sockets) or when the communication between subassemblies must be
tested (such as with a backplane or daughter card situation). This is particularly so
when signaling at high speeds, but as we will see in later chapters, signal integrity
problems are not limited to high-speed circuits.
In any of these cases it can be more cost effective to design and build circuit
boards with limited functionality specifically designed to test the high-risk portions of the actual design. These printed circuit boards are called test boards to
distinguish them from evaluation boards. As discussed in Section 4.3, evaluation
boards are often provided to customers for evaluation of one or more ASICs in a
well-defined environment.
4.2
Test Boards
The signal integrity engineer should consider test boards as opportunities to reduce
risk and to further their own understanding of the way simulation portrays physical effects.
For instance, laboratory measurements can be made on the test board to:
•
Validate I/O and connector circuit simulation models.
•
Validate a routing strategy in a high-density pin or connector field.
•
Validate a stackup.
•
Validate the performance of a split plane.
•
Test power supply decoupling strategies, including component values and
types, and layout rules.
31
32
Printed Test and Evaluation Boards
•
Validate circuit models of various routing topologies, and to decide between
them.
•
Determine if a given trace width has adequately low loss and compare to the
simulation.
•
Determine crosstalk of critical routes, especially in dual stripline situations
that are difficult to simulate properly.
•
Validate I/O cell circuit models.
•
Determine the ability of a manufacturer to meet impedance specifications.
•
Determine the characteristics of traces routed near the board edge or other
cutouts in the board.
However, it is important for these boards not to become too intricate, especially if the experiments require excessive support from others. For instance, a dual
stripline crosstalk test would require that many lines be driven and some number
of them monitored. Support from other engineers will be required unless the signal
integrity engineer can design the logic or create the firmware necessary to do this.
In fact, it is critical not to lose sight that test boards are intended to reduce risk and
not to create it by siphoning off engineering resources that might be better used in
the development of the actual product.
4.2.1 Test Boards as a Chip Debug Platform
An important use of test boards is as a platform to facilitate the test and debug of
a new ASIC. For instance, a fabless semiconductor company would create such a
board so their engineers can validate and demonstrate the features of their chip in
an electrically controlled environment. This can be particularly helpful when providing customer support if the applications engineer can configure the ASIC in the
same way as the customer and demonstrate the same problem.
A test board is an important engineering tool because it is far less costly to debug a new chip design on a test card than it is on the manufacturer’s automatic test
equipment (ATE). Although deep, high-quality test patterns running on ATE are
critical for the manufacturer to determine if the ASIC works the way the designers
had intended, tracking an elusive bug on the ATE is inefficient and costly.
Instead, a test board can be developed that provides an electrically sound environment for the ASIC that includes jumpers, LEDS, test points, and connectors
that allow the designer to configure and test the chip under normal and extreme
conditions.
Test boards often have connections to visibility points that allow the designer
to observe or set logic deep within the ASIC. This is particularly helpful when the
chip can be placed in one or more test modes. Such modes are often not intended
for customer use and frequently redefine the normal function of ASIC I/O pins.
With a properly designed test card, the designer can use these pins in a test mode
without interference from the logic to which they would normally be connected.
A well-coordinated debug strategy includes a test plan that describes how to
use the test card with whatever test logic has been designed into the chip. In this
way test patterns can be written and any special fixtures or long lead time test
4.3
Evaluation Boards
33
equipment and other parts can all be in place before the test card arrives in the
laboratory.
4.2.2 Margining Power Supply
With the proper laboratory equipment, the power supply to the chip can be varied
(margined) to see how low or high the voltage can be set before the chip begins to
malfunction. Margining can also be used to determine which circuits are the first
to fail. This technique is especially powerful when the temperature of the air surrounding the integrated circuit is simultaneously margined. As shown in Chapter
2, in general, low voltage and high temperature represent the extreme slow case
for CMOS integrated circuits, while low temperature and high voltage generally
represent the fast operation, high-current draw case.
4.2.3 Testing Decoupling Capacitor Efficacy
The frequency content of the current drawn by an integrated circuit or a group of
chips under various switching and operating conditions can be determined by measuring the current and performing a Fourier analysis on the waveform to obtain the
frequency spectra. Many modern oscilloscopes will directly perform a fast Fourier
transform (FFT), or the readings can be saved and externally processed by mathematical CAD programs such as MathCAD [1] or MATLAB [2].
By observing the spectra under various conditions, including with various I/O
switching patterns, the signal integrity engineer can determine the frequencies over
which the power supply decoupling network is operating. Specific frequencies with
particularly high amplitudes can be identified, and if necessary, decoupling can be
specifically designed to eliminate them.
4.2.4 Testing Sensitivity to Clock Quality
The chips sensitivity to clock amplitude, frequency, duty cycle and jitter can be
tested by replacing the expected clock source (often a crystal oscillator or PLL/
fanout circuit) with a laboratory function generator or test set. In this way the
sensitivity of the chip to clock quality, especially when tested over extremes of temperature and power supply voltage, can be determined.
4.3
Evaluation Boards
Ideally, the test board used for debug can be given to prospective customers for
them to evaluate the operation of the ASIC, but in practice it is usually better to
have a different board for this purpose. In the past, the strategy of creating a single
test and evaluation board made economic sense when circuit board artwork could
only be created by layout specialists and when fabricating and assembly shops were
reluctant to take on small sized orders.
While it is true that the routing of high-speed, high-performance circuit boards
is best left to experienced specialists, the proliferation of layout design CAD tools
(some of which include auto routing software that is driven from the schematic to
34
Printed Test and Evaluation Boards
route traces automatically) makes it possible for the skilled signal integrity engineer
to lay out simple, small, multilayer circuit boards for customer use.
Often these “eval boards” are small and contain only the minimum amount of
peripheral circuitry necessary to demonstrate operation of the ASIC. Connectors
are provided as necessary so that the user can connect the evaluation board to his
or her application.
Although the interconnect present on these boards is often simple, careful
thought must be given to the stackup and to decoupling and filtering. These must
provide the electrical environment necessary to demonstrate the best possible operation of the ASIC. However, unnecessarily complex decoupling, or a complex
stackup (for instance, using many power and ground planes or planes that have
multiple splits) can be seen by the potential customer as an undesirable requirement for their design. In fact, often the customer will request (and sometimes will
require) copies of the evaluation board artwork files so that the layout of the end
product will exactly match the evaluation board. In this way the ASIC will operate
in the same electrical environment in the end product and so should behave in the
same way. An overly complex stackup in the evaluation board can lead to higher
costs in the end-user board, which could discourage potential customers.
4.4 Roll of the Printed Circuit Layout Shop
A layout designer is the specialist who creates the artwork files required by a circuit
board manufacturing shop to actually build the circuit boards. They are employed
in firms ranging in size from a single proprietor operation to a multiperson office to
a company that has staff located throughout the world.
Layout designers understand the manufacturing requirements and technological limitations imposed by manufacturers, but usually they are not expert in circuit
design or signal integrity. Instead, in the absence of specific rules specified by the
customer, they create the artwork by using previous experience of apparently similar designs and general rules of thumb.
For this reason it is important that the designer or firm has experience laying
out boards in the same technological field as the test or evaluation board. For instance, a firm that customarily designs back planes and high-end server boards may
not have meaningful experience in the automotive or medical electronics fields. In a
similar way, a layout designer experienced with aerospace electronics circuit boards
may not be the best choice to design a low-cost ASIC evaluation card.
Although the layout shop may not have direct experience in signal integrity,
often they work closely with signal integrity experts, and commonly the large layout firms have signal integrity specialists on staff or under contract. This can be a
helpful resource when designing evaluation and test cards, but in general the layout
designer will not seek help from these specialists without prior approval.
It is customary for the layout shop to send the incomplete layout database at
regular intervals to the customers so they can track progress and to inspect the
artwork. Usually the project manager and the circuit designer will review these
snapshots for schedule and design compliance, but the signal integrity engineer
should also examine them to ensure that the layout is in compliance with whatever
4.5 Fabricating and Assembling the Board
35
signal integrity rules have been specified and that it in general conforms to good
engineering practice.
4.5 Fabricating and Assembling the Board
The printed circuit fabrication shop (the fab shop) is the manufacture that uses the
circuit board artwork designed by the layout shop to create the actual printed circuit board. Generally, the layout design shop will directly send the layout artwork
files to the board shop selected by the customer. Often the layout shop can recommend a fab shop with which it has worked that is most suitable for a particular
design.
The boards delivered from the fab shop are drilled and punched to accept components such as connectors and some electronic components, but the fab shop does
not install components. That task is the responsibility of an assembly shop or of a
contract manufacturer.
If requested, for controlled impedance boards the fab shop will supply the customer with the test coupons they used to verify that they had achieved the target
impedance value. If the customer has access to a time domain reflectometry (TDR),
these coupons can be helpful in determining the actual impedance of a given lot of
boards. If requested, the fab shop can also provide test reports showing the trace
impedance and line widths.
Like layout firms and circuit board manufacturers, contract manufacturers
range in size and in capabilities. Some specialize in specific markets (such as aerospace, automotive, or telecommunications), but the larger firms tend to work in
all markets. The largest contract manufacturers provide design and consulting services, which can include schematic capture, signal integrity analysis, circuit board
artwork generation, circuit board fabrication, circuit board assembly (attaching
components to the circuit board), and electrical testing once all the components
have been attached (“final test”).
4.6 Alternative Methods for Design and Fabrication
Using an experienced layout designer specialist and a circuit board fab shop with
moderate to high-end processing capabilities remains the lowest-risk way to produce complex or high-performance test and evaluation boards. However, the creation of small test and evaluation circuit boards is now straightforward with the
proliferation of easy-to-use circuit board layout software and the increasing number of fab shops willing to manufacture circuit boards in prototype quantities. This
provides an opportunity for the signal integrity engineer to design these boards
without using a specialty layout shop. Small start-up companies and small fabless
semiconductor firms often find this approach particularly attractive.
The cost per board will be reasonably low if the boards have only a few layers
(usually 4 or less) and if the board area is no more than roughly 25 square inches
(160 square cm). The signal integrity engineer can still lay out and have manufactured boards that are larger than this and that have more layers, but the price of
the CAD software necessary to layout these boards will increase, and fewer shops
36
Printed Test and Evaluation Boards
will be willing to manufacture the board at reasonable cost in small (prototype)
quantities.
Although prototype fab shops routinely manufacture circuit boards suitable
for surface mount technology (in some cases including BGA technology), usually
the lowest-cost shops do not support the most advanced technologies. In general,
the low-end shops will require larger line width and wiring pitch design rules than
higher-cost shops, and impedance control may not be available as an option.
It is common for a low-end fab shop to specify the available drill and via sizes
(including the pad size) from which to choose. In virtually all cases the antipad sizes
are automatically determined by the via and pad size, and cannot be changed. This
means that by using these shops the signal integrity engineer is constrained to using
only those specific sizes and custom via/pad/antipad design intended to produce a
specific impedance (discussed in Chapter 14) cannot be designed.
Before selecting a fab shop to make a test or evaluation board, the signal integrity engineer should determine the trace width and pitch requirements for the
board. Since not all shops have the same capabilities, selecting the width and pitch
will eliminate many shops form consideration, especially if traces and spaces narrower than 5 mils (0.13 mm) are required. Many shops now routinely etch traces
with a width of 4 mils (0.11 mm) or less, but 5 mils remain the most narrow that
the least expensive shops will fabricate.
4.7
Main Points
•
Test boards are a good way to engineering debug and evaluate custom integrated circuits and signal integrity structures.
•
Evaluation boards for customer use are an effective way to demonstrate new
products.
•
Moderately complex test and evaluation boards can be created by the signal
integrity engineer without using layout specialists.
References
[1]
[2]
Parametric Technology Corp., Needham, MA, http://www.ptc.com.
The MathWorks, Inc., Natick, MA, http://www.mathworks.com.
CHAPTER 5
Printed Circuit Board Construction
5.1
Introduction
Understanding circuit board construction and terminology helps the signal integrity
engineer discuss trade-offs with board designers and board fabrication shops. Understanding the materials used to form circuit boards (and the way in which they
are specified) is crucial when performing signal integrity analysis.
In this chapter we describe characteristics of circuit boards common to all multilayer circuit boards (even though the construction details often differ between
fabrication shops), and the terms commonly used in the circuit board industry are
introduced and defined.
5.2 How Is a Multilayer Circuit Board Constructed?
Multilayer circuit boards are made by applying heat and pressure to a stackup
of fully cured laminate cores and partially cured prepreg mats. Circuit traces are
photolithographically patterned and then etched on the copper sheets. A six-layer
board stackup is shown in Figure 5.1.
The prepreg is a mat of woven glass fibers saturated with partially cured resin.
FR4 grade epoxy resin is most common, but other chemistries are available. Pressing together the stackup under heat fully cures the prepreg. In contrast, a core is a
fully cured prepreg mat that has copper sheets (foil) bonded to one or both sides.
The heat and pressure cause the soft prepreg to flow and form around the copper traces patterned on the core. Once cured, the prepreg hardens and forms a tight
bond between cores, securing them.
Manufacturers can create the circuit board outer layers on a core (core construction as shown in Figure 5.1) or on prepreg (foil construction as illustrated
in Figure 5.2). Because it offers better control and requires less processing, many
manufacturers prefer to use foil construction [1].
5.2.1 How Are Connections Made Between Layers?
Holes drilled through the circuit board dielectric materials form tube-like structures
called vias (more descriptively, plated-through holes). Although the process varies,
37
38
Printed Circuit Board Construction
Exposed microstrip
(”outer layer”)
Prepreg
mats
Ground plane
attached to laminate
Laminate
Stripline (”inner layer”)
attached to laminate
Figure 5.1 A six-layer stackup using core construction. Copper traces or planes are shown in black,
and cores are shown in white. Prepreg mats are shaded.
PTH
Blind
Buried
L1
Copper
clad core
L2
L3
L4
L5
L6
Prepreg
Antipad
Figure 5.2 Circuit board vias in a six-layer foil construction board. The PTH connects L1 and L4.
Note the nonfunctional pad on L3. The buried via connects L3 to L4. The blind via connects layer
L1 to L2.
commonly the plating process deposits copper inside the tube, making the platedthrough hole (PTH) a conductor.
The PTHs have circular pads (rings of copper) that connect the via to signal
traces or power/ground planes. Three types of vias commonly found on circuit
boards are shown in Figure 5.2.
This stackup has six copper layers (L1 through L6). Signals are routed on layers L1, L3, L4, and L6. Layer L2 is a ground plane, while L5 is a power plane.
The PTH passes completely through the board but only connects L1 to L4.
5.3 What Is the Significance of Calling a Circuit Board “FR4”?
39
Antipads (holes precisely created in the copper planes) allow the PTH to pass
through (L2 and L5 in the figure) without connecting.
The nonfunctional pad on L3 does not provide an electrical connection to traces on L3, but it does help anchor the PTH in the stackup. The parasitic electrical
effects of these pads can be detrimental to very high-speed signals. Depending on
the board thickness, high-end fab shops can selectively remove nonfunctional pads
on selected vias.
Blind vias (shown connecting L1 to L2) are PTHs that start on either of the
outer layers but do not pass entirely through the board. Blind vias can connect to
more than just the two layers suggested in the figure.
The buried via (sometimes mistakenly called a blind via) is shown connecting
together signals on L3 and L4. It is a PTH that does not pass completely through
the board and is not visible from the outer layers.
5.2.2 What Is the Via Aspect Ratio?
A via that is much longer than its outer diameter can be unreliable and is difficult
to manufacture [2]. The via aspect ratio calculated in (5.1) is a way to gauge this
difficulty:
Via aspect ratio =
PWB _ thickness
O.D.
(5.1)
Some shops call the via aspect ratio the drilling ratio, or simply the aspect ratio.
A via aspect ratio of 5 means that the fab shop can reliably produce boards that
can be five times thicker than the outer diameter (O.D.) of the smallest via (the pad
size is not a factor). For a 10-mil diameter via, this means that the board can be no
more than 50 mils thick.
Presently, boards with aspect ratios of 5 or less are considered as having mainstream volume production worldwide. Aspect ratios in the 6–8 range are usually
considered standard production, ratios in the 9–14 range are considered advanced
or low volume, and ratios greater than about 14 are considered prototype or
experimental.
The signal integrity engineer uses the via aspect ratio provided by the manufacturer and the estimated board thickness to determine the minimum sized via that
can be used in a design (for example, when escaping from a small pitch BGA).
5.3 What Is the Significance of Calling a Circuit Board “FR4”?
Although often used to describe the electrical characteristics of a typical circuit
board, the term FR4 has little electrical significance. It actually refers to the National Manufacturers Association (NEMA) level of fire retardation.
Ratings FR1 through FR3 are paper-reinforced epoxy resin systems, while ratings FR4 and 5 are glass fabric–reinforced epoxy resin systems with “reduced burning rate” [3].
40
Printed Circuit Board Construction
Table 5.1 shows two important electrical parameters for a small number of the
FR4 class laminate systems available to fabrication shops. As discussed in Chapter
7, Dk refers to the dielectric constant, and LT refers to the loss tangent.
The dielectric constant is different for each vendor’s version of FR4, and all the
values are lower than the 5.4 worst-case ASTM specification. Because trace width,
impedance, and stackup thickness are all related to the dielectric constant, Table
5.1 illustrates why different fabrication shops will produce FR4 boards having very
different electrical characteristics.
5.3.1 Why Is There Such Variability in Dk Between Laminate Manufacturers?
A material called E glass is most often used for the reinforcing fibers, but some
higher-performance laminates use S glass or NE glass because of their lower Dk
and LT values.
Various resins can be used to bind the fibers. As indicated in Table 5.2, the
glasses and resins do not have the same electrical characteristics. In fact, the relative
amount of glass fibers to resin (the glass-to-resin ratio) determines the Dk and LT
values for a fabricated core or laminate mat.
The relationship between core thickness and resin content for FR4 with E glass
as defined in IPC-CC-110 [9] is illustrated in Figure 5.3.
Industry has not universally agreed on the glass-to-resin ratio to use when publishing Dk values. This lack of standardization complicates comparing data sheet
values; the comparison is only valid when the resin content (and the glass weave)
is identical.
Table 5.1 Electrical Characteristics for a Sampling of FR4
Laminate Systems
Dk
LT
Dk
LT
Name
(100 MHz)
(1 GHz)
Reference
ASTM D 1867
5.4
0.035
—
—
[3]
FR402
4.6
0.016
4.25
0.015
[4]
PCL226
4.5
0.015
4.3
0.015
[5]
IS410
4.0
0.015
3.9
0.019
[6]
Table 5.2 Electrical Properties at 1 MHz
of Circuit Board Glass and Resins [7, 8]
Material
Dk
LT
E glass
6.2
0.004
S glass
5.2
0.003
NE glass
4.4
0.0012
FR4 epoxy resin
3.6
0.032
BT resin
3.1
0.003
Polyamide resin
3.2
0.02
Cyanate ester resin 2.8
0.002
5.3 What Is the Significance of Calling a Circuit Board “FR4”?
41
4.8
4.7
Dk
4.6
4.5
4.4
4.3
4.2
4.1
40
45
50
55
60
65
70
75
% Resin Content
Figure 5.3 FR4 dielectric constant depends on the glass-to-resin ratio. Because the resin has a much
lower Dk than E glass, Dk falls as the amount of resin increases. Raw data from [9].
5.3.2 Why Is the Same Board Produced Differently by Various Fab Shops?
Sending the same artwork to multiple circuit board fabricators does not guarantee
that each shop will produce boards having the same electrical characteristics. In
fact, unless the board is very simple, it is nearly certain each shop will use a unique
combination of prepreg and core thicknesses.
For example, one shop might choose to use two thick prepregs with a 42%
resin content to form a layer, while another shop might use four thinner sheets
each having a 53% resin content to obtain the same overall layer thickness. As illustrated in Figure 5.3, this difference in resin content changes the Dk from 4.65 to
4.35. As explored in later chapters, this difference affects the signal time of flight
and the trace impedance (and possibly the trace width).
5.3.3 What Are Some Alternatives to FR4?
Manufacturers have created a variety of laminate systems, which are electrically
and mechanically (but sometimes only mechanically) superior to common FR4
resin systems. Table 5.3 shows a specific product offering from various vendors for
the chemistries appearing in Table 5.2.
Megtron 6 is a PPO (polyphynoleneoxide) chemistry offered by Panasonic
[10]. Generally, the Dk and LT values for PPO-based laminates are less affected by
frequency than standard FR4 systems. G200 is a BT (bismaleimide/triazine blend)–
based chemistry from Isola [11]. The higher temperature capabilities and lower
moisture uptake make these laminates attractive to the micropackaging industry.
Nelco N7000-1 [12] is a polyimide-based system. Polyamide circuit boards can
withstand high temperatures and are often used in commercial test equipment and
in aerospace applications. Nelco N8000 is a cyanate ester [13] system offered by
Park Electrochemical Corporation. Laminates based on these chemistries are often
42
Printed Circuit Board Construction
Table 5.3
Chemistry
A Sampling of Alternate Laminate Systems
Product Name Dk/LT at 10 GHz
PPO
Megtron 6
3.4/0.004
BT
G200
3.65/0.015
Polyamide
N7000-1
3.8/0.016
Cyanate Ester N8000
3.5/0.011
Ceramic
3.4/0.0022
RO4003
used in RF applications. The RO4000 series of materials from the Rogers Corporation [14] are reinforced hydrocarbon/ceramics that have low and stable Dk values.
Theses are popular in high-performance digital and RF applications.
5.4 Circuit Board Traces
The copper conductors used to interconnect circuits (the traces) are formed either
on the boards surface (the outer layers) or buried within the circuit board material
(inner layers). Outer layer traces are called microstrip (or microstripline); inner
layer traces are called stripline.
The four basic topologies used to interconnect digital circuits are shown in Figure 5.4. Microwave and RF designers use additional topologies that are not shown.
The outer layer traces can either be exposed microstrip or embedded microstrip,
and the inner layers can be either stripline or dual stripline.
Soldermask
Trace
Dielectric
Plane
Exposed microstrip
Stripline
Embedded
microstrip
Dual
stripline
Figure 5.4 Types of circuit traces used to connect digital circuits. Microstrips are formed on the
outer layers, while inner layers create striplines.
5.4 Circuit Board Traces
43
5.4.1 Microstrip
Exposed microstrip is formed when a copper trace is separated from a metal plane
by a block of dielectric material. The trace forms a transmission line with the plane
(the return path), which is connected to either power or ground. In exposed microstrip the copper trace is directly exposed to the environment. It is not covered by
an encapsulate or other nonconducting sealant materials.
This contrasts with embedded microstrip, which is a microstrip covered by an
insulating material such as solder mask. On commercial digital logic circuit boards,
virtually all microstrips are covered with solder mask, making them embedded.
Although commonly called microstrips, Figure 5.4 shows that they are actually
embedded microstrips.
Solder mask is an insulating material used to prevent solder bridges forming
between traces. Coating the traces with this dielectric material lowers the trace
impedance and increases the trace capacitance and delay time. To obtain a specific
impedance, the manufacturer will adjust the trace parameters (generally by making
the trace less wide) to compensate for the solder mask.
Microstrips are made thicker than the base copper layer as part of the plating process used to create vias. The amount of plating necessary is determined by
factors associated with forming the via, and the final microstrip thickness is not
usually the primary consideration for the manufacturer. An example appears in
Figure 5.5.
5.4.2 Stripline
Stripline traces are sandwiched between two metal planes. A block of dielectric
material separates the planes and trace. Often striplines are thought of as being
centered between the planes, but usually this is not the case in practice. Instead,
because of the way in which circuit boards are manufactured, offset stripline (where
the trace is located closer to one of the planes) is the norm.
It is possible to sandwich two traces between the planes rather than just one.
Such dual stripline increases routing density, but can produce excessive noise coupling (crosstalk, discussed in Chapter 10) between the traces.
By routing the traces parallel to and directly underneath one another, the coupling can be intentionally accentuated, and dual stripline can form a broadside
coupled differential pair (discussed in Chapter 13).
5.4.3 What Is a Mil and What Does a Trace Thickness in Ounces Mean?
It is common practice in North America to define the copper trace width in mils and
the thickness in terms of copper weight, which is measured in ounces per square
foot. A mil is 0.001 inch and is equal to 25.4 μm (0.0254 mm).
Common copper thickness values are 1/2-, 1-, and 2-ounce copper, but others
are available. Table 5.4 shows the minimum copper thickness in mils and μm required by IPC-2221 [15] for inner and outer layer traces.
Table 5.4 shows that a half-ounce trace is nominally 0.67-mil (17 μm) thick. It
must be at least 0.47-mil (12 μm) thick when on an outer layer (such as microstrip)
and 1.3-mils (33 μm) thick when on an inner layer (such as stripline).
44
Printed Circuit Board Construction
Plating
Foil
Figure 5.5 Outer layer trace is made thicker than the base metal by plating. (Courtesy of TTM
Technologies, Inc. Used with permission.)
Table 5.4 Minimum Copper Thickness by Weight Per
IPC-2221 [15]
Nominal
Copper Foil Thickness Internal Layer External Layer
Weight
Mils (μm) Mils (μm)
Mils (μm)
1/8
1/4
0.35 (9)
3/8
0.14 (3.5)
0.79 (20)
0.24 (6)
0.87 (22)
0.31 (8)
0.98 (25)
1/2
0.67 (17)
0.47 (12)
1.30 (33)
1
1.38 (35)
0.98 (25)
1.81 (46)
2
2.76 (70)
2.20 (56)
3.00 (76)
3
3.58 (91)
4.21 (107)
4
4.80 (122)
5.39 (137)
Not all weights have the nominal thicknesses specified.
Although a maximum thickness is not defined, to manage material costs and
to improve etching, the internal layer copper is usually not much thicker than the
IPC nominal values.
Typically after processing, the half-ounce copper used in internal layers are 0.6
± 0.1-mils thick and are 1.24 ± 0.25-mils thick for 1-ounce copper [16].
From experience, outer layer half-ounce copper thickness ranges from 1.8 mils
to 2.7 mils, depending on the board thickness. Other studies [16] have found the
average to be between 2.1 mils and about 2.6 mils. That same study showed that
1-ounce copper had an even greater variation, ranging from 1.3 to 3.5 mils, depending on the vendor.
5.4 Circuit Board Traces
45
5.4.4 What Copper Types Are Used to Form Traces and What Are Their
Characteristics?
The copper foil used on circuit boards is created by either an electrodeposition or
rolling process. The end result from either process is very pure copper foil, but the
foil’s mechanical and high-frequency electrical characteristics are quite different.
Electrodeposited (ED) foils are formed on a drum rotating in a solution containing copper [3]. This leaves one side of the foil smooth and the other side rough.
The rough side of these foils attach to the laminate. This foil is the most commonly
used in commercial circuit boards. Double treat foils roughen both sides. Reverse
treat foils roughen the normally smooth side of the foil. With these foils the smooth
side attaches to the laminate. These configurations are illustrated in Figure 5.6.
In contrast, rolled (also called wrought) foils are created by squeezing copper sheets through closely spaced rollers. This hardens the copper and crushes the
copper crystals, making the copper smooth on both sides. In fact, rolled foils are
smoother and denser than ED foils. For this reason the bulk resistivity of rolled
foils is lower than that of ED foils [17], as is apparent in Table 5.5. In later chapters
this information is used to determine the DC resistance and AC loss of traces.
Although the majority of ridged circuit boards for the commercial market use
ED foils, rolled foils are often used in flex circuitry [10] and in those situations
where there is a large temperature swing (such as space and avionics applications)
[17]. As shown in Chapter 8, high-frequency loss depends in part on the smoothness of the copper surface. For this reason, rolled foils are often used in RF work.
5.4.5 What Is Surface Roughness and What Are Typical Values?
To promote adhesion, the copper foil is mechanically roughened and various adhesion, barrier, and stabilization layers are added [19]. These steps are necessary
Copper
trace
Laminate
Typical
Double
treat
Reverse
treat
Rolled
Figure 5.6 Roughness of various copper foils, exaggerated for clarity. Typical foils are rougher on
the laminate side, while double treat is rough on both sides. Reverse treat attaches on the smooth
side, and rolled copper is smooth on both sides.
Table 5.5 Resistivity in μΩ-cm for
Copper Traces [18]
Weight
Ounces (μm) ED
Rolled
1/2
1.82
1.78
1
1.78
1.74
2
1.78
1.74
46
Printed Circuit Board Construction
because a trace (especially if it is narrow) would easily peel away from the substrate
without peaks and valleys in the copper to grab hold and anchor it to the dielectric.
To improve peel strength on commercial circuit boards, ED foils with a more
pronounced tooth profile are usually used on outer layer traces, while smoother
ED foils are usually used on inner layers [20]. However, to promote adhesion with
the prepreg, the top surface of inner layer copper is sometimes “micro-roughened”
[21] as well.
Surface roughness is specified by foil manufactures in terms of the maximum
valley to peak height of the ridges, or as Ra, their RMS average sampled over a
region. Their amplitude and shape depend on how the copper is formed and on the
surface treatments used to further roughen the surface.
As shown in Table 5.6, average surface roughness values for electrodeposited
copper range from 0.7-μm RMS to as much as 3-μm RMS [18–22]. This wide
range reflects the diversity of copper foil types and roughening techniques used in
the industry. Copper foil adheres better to some laminates than others, so the copper used with alternate laminate systems is sometimes made rougher than that used
in conventional FR4 systems [9].
5.4.6 Are Traces Rectangular?
Etched traces are not usually rectangular in shape because the etching solution attacks the copper both vertically and horizontally. This results in the traces being
more trapezoidal than rectangular. This was first shown in Figure 5.5 for an outer
layer trace, but as shown in Figure 5.7, the effect also happens on inner layer traces.
Figures 5.5 and 5.7 show the trace as wider at the bottom than the top, but it is
possible for the trace to be wider at the top than the bottom. Because etching times
are shorter, thinner copper generally tends to produce more rectangular traces than
thicker copper. Similarly, rough copper requires more etching time than smooth
copper, so smooth copper traces tend to be more rectangular than rougher traces.
5.5
Main Points
•
Multilayer circuit boards are created by fusing together prepreg mats and
laminate cores.
•
The glass-to-resin ratio is an important factor in determining trace impedance and delay.
•
Microstrips are formed on the circuit board outer layers.
Table 5.6 Typical Average Roughness Values
for Treated Copper Foil [18–22]
Copper Weight Ra ED
Ra Rolled
(Ounces)
(μm, RMS) (μm, RMS)
1/2
0.7–1.9
0.3–1.4
1
1.0–2.4
1.2–1.4
2
1.1–3.0
1.4
Problems
47
Smooth
upper
surface
Rough
lower
surface
Figure 5.7 Cross-section of a 1-ounce trace. Etching causes traces to be trapezoidal in shape. The
roughness of the copper surface is also apparent. (Courtesy of TTM Technologies, Inc. Used with
permission.)
•
Striplines are formed on inner layers.
•
The traces are generally trapezoidal rather than rectangular in shape.
•
FR4 is a level of fire retardant and does not define the electrical characteristics of a board system.
•
Advanced laminate systems can use resins other than the traditional FR4
systems.
•
The plating process creating vias makes the shape and thickness of microstrip
traces less predictable than stripline traces.
•
In North America, trace thickness is specified in terms of copper weight in
ounces.
•
Traces are usually formed with electrodeposited copper foils. Generally, these
foils are rougher than the rolled foils used in microwave work.
Problems
Answers to these problems are available on the Artech House Web site at http://
www.artechhouse.com/static/reslib/thierauf/thierauf1.html.
5.1
5.2
5.3
5.4
The signal integrity engineer reviewing your completed FR4 circuit board
informs you that the conductor loss is too high on some high-speed nets.
What manufacturing options should you consider?
In reviewing signal integrity models, you notice that the half-ounce stripline
and microstrip traces were assumed to be 0.65-mil thick. Is this valid?
Prototype quantities of your circuit board have been manufactured by a
high-end production shop, but once production begins, the design will be
transferred to a manufacturer more accustomed to high-volume production. What are the signal integrity risks?
Given the scenario described in Problem 5.3, what signal integrity tests
should be performed on the volume production boards?
48
Printed Circuit Board Construction
5.5
Based only on the via aspect ratio, what is the smallest sized through-hole
via that you should expect in standard production if your circuit board is
0.092-inches (2.3 mm) thick?
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
Merix Corp., Design for Manufacturability of Rigid Multi-Layer Boards, Revision 7/99,
Forest Grove, OR, July 1999.
Goval, D., “Reliability of High Aspect Ratio Plated Through Holes (PTH) for Advanced
Printed Circuit Board (PCB) Packages,” Annual Proceedings, IEEE International Reliability Physics Symposium, Denver, CO, April 8–10, 1997, pp. 129–135.
Shugg, W. T., Handbook of Electrical and Electronic Insulating Materials, 2nd ed., New
York: IEEE Press, 1995, pp. 270–271.
“Isola FR402,” Isola Group, Chandler, AZ, data sheet #5023/5/02, Rev 3/16/07.
Polyclad Laminates, Inc.,” PCL-FR-226/PCL-FRP-226,” data sheet PCL-226, Isola Group,
Chandler AZ, October 2001.
“Isola IS410,” Isola Group, Chandler, AZ, data sheet IS410 08-13-08.
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Coombs, C. F., Jr., Printed Circuits Handbook, 6th ed., New York, McGraw-Hill, 2008.
Jawitz, M. W., Printed Circuit Board Materials Handbook, New York, 1997.
“DATA SHEET Megtron 6,” Panasonic PCB Materials, Santa Ana, CA, undated.
“Isola G200,” Isola Group, Chandler, AZ, data sheet, Rev 5-16-08, dated 2006.
“Nelco® 7000-1,” Nelco Advanced Circuitry Materials, Park Electrochemical Corp., data
sheet Rev 2-07 (undated).
“Nelco® 8000, Nelco N8000Q,” Nelco Advanced Circuitry Materials, Park Electrochemical Corp., data sheet Rev 2-07 (undated).
Rogers Corp., “RO4000 Series High Frequency Circuit Material Data Sheet,” No. 92-004,
Advanced Circuit Materials, Chandler, AZ, 1992.
Institute for Interconnection and Packaging, Electronic Circuits Generic Standards on
Printed Board Design, IPC-2221 2215, Northbrook, IL, February 1998.
Brist, G., “Design Optimization of Single-Ended and Differential Impedance PCB Transmission Lines,” PCB West Conference Proceedings, 2004.
Merix Corporation, “Electrodeposited vs. Rolled Copper,” Merix Corporation, www.
merix.com.
Rogers Corp., “Copper Foils for Microwave Circuits,” Publication #92-243, dated 2/00,
Microwave Materials Division, Chandler, AZ.
Hilburn, R., et al., “The State of Copper,” Printed Circuit Design & Manufacturer, May
2005, pp. 24–29.
Brist, G. et al., “Non-Classical Conductor Losses Due to Copper Foil Roughness and Treatment,” ECWC 10/IPC/APEX Conf., Fall 2005.
Lee, B., “The Impact of Inner Layer Copper Foil Roughness on Signal Integrity,” Printed
Circuit Design & Manufacturing, May 2007, pp. 27–29.
Normyle, S. J., et al., “The Impact of Conductor Surface Profile (Rrms) on Total Circuit Attenuation in Microstrip and Stripline Transmission Lines,” White Paper, Taconic Advanced
Dielectric Division, Petersburg, NY, undated.
CHAPTER 6
Transmission Line Fundamentals
6.1
Introduction
In this chapter we use a circuit analysis approach to introduce fundamental transmission line concepts. This approach is more intuitive than field theory, but [1–3]
should be helpful for those interested in exploring transmission lines in that way.
We begin by defining a transmission line and its return path and developing a
transmission line circuit model. We will restrict our study here to transmission lines
having low losses, because this lets us develop an intuitive understanding without
the mathematics necessary when losses cannot be ignored. Lossy transmission lines
are discussed in Chapter 8. We discuss return paths in some detail because return
path management is an important aspect of signal integrity analysis. We conclude
the chapter by developing rules of thumb that define when a conductor acts as a
simple lumped circuit or when the circuit elements are distributed.
6.2 What Is a Transmission Line?
As illustrated in Figure 6.1, a transmission line is formed with two or more conductors. The signal conductor carries the signal energy from a generator to a load,
and the second conductor (the return) completes the circuit by returning the signal
current back to the generator. The signal conductor could be a microstrip trace and
the return path the underlying ground plane. The generator is any signal launching
device, such as an ASIC I/O driver.
6.2.1 What Is the Signal Return Path?
Frequently, signal integrity engineers speak of the “signal return path,” but often
it is taken for granted and not fully understood. Studying the return path is important because it determines the trace’s capacitance, inductance, resistance, and noise
coupling.
Because capacitance is often more intuitive than inductance, we will ignore
inductance for now and focus on the importance of the return path in determining the capacitance. Resistance is also affected by the return path, but we will wait
until Chapters 7 and 8 to discuss this important topic.
49
50
Transmission Line Fundamentals
Return
Load
Copper
plane
Signal
conductor
Dielectric
sheet
Microstrip
trace
Return current
Signal current
Generator
Figure 6.1 A transmission line created by a signal trace and its return.
The side view of a trace over a ground plane is shown in Figure 6.2. We can see
that the trace is one plate of a capacitor with the second plate being the return path
(such as a ground plane). The distributed inductance is formed by the current loop
area between the trace and ground plane. The capacitance and inductance values
are determined by the width of the trace and the distance to the ground plane.
As we will see, the inductance and capacitance values determine the trace impedance. This is why an impedance value is really only meaningful when it is quoted with respect to a specific return plane. We will return to this later in this chapter
and somewhat more indirectly in Chapter 7.
6.3 Circuit Model of a Transmission Line
A circuit model for a small segment of trace is shown in Figure 6.3. The trace has
a resistance (R) and an inductance (L). The capacitance formed between the trace
and ground plane is modeled by capacitor C, and the capacitor losses are modeled
Figure 6.2 Distributed parallel plate capacitors and loop inductance formed between the signal
trace and the ground plane return path.
6.4
Impedance and Delay
51
by conductance G. In Chapter 7 we discuss typical values for these parameters, and
in Chapter 17 we show how to calculate them. We will avoid this level of detail
here and note that the entire trace is made up of an infinite number of identical
copies of these tiny segments. In fact, the segments are so small they blend together
smoothly. The resistors, inductors, and capacitors are distributed evenly along the
length of line.
The circuit in Figure 6.3 is a constant-k lowpass filter, and circuit theory can be
used to analyze its behavior. Taking this approach lets us develop a circuits-based
intuitive understanding, but in the following chapters we will use some concepts
from field theory when explaining the different characteristics of various types of
circuit board transmission lines.
The circuit has the following characteristics:
•
It imposes a delay. Time is required for the signal to travel from one end of
the line to the other. This is called the transmission lines propagation delay
time, tpd.
•
The line possesses an impedance. The current launched down a such a line is
limited by the transmission lines characteristic impedance, Zo.
•
The line exhibits loss. The signal is attenuated and, if the loss is severe enough,
the signal will become distorted. This is discussed in Chapter 8.
6.4 Impedance and Delay
By using circuit analysis [4–6] and ignoring the resistance and conductance (R and
G), the characteristic impedance and the delay of a long chain of RLCG circuits
such as Figure 6.3 are found to be:
Zo =
L
C
tpd = LC
(6.1)
(6.2)
Impedance (Zo) has units of ohms and it has that value for any length of line.
For instance, a transmission line may have an impedance of 75Ω, but it does not
R
L
G
C
Figure 6.3 Trace circuit model. The trace’s distributed circuit elements are represented by a string
of closely spaced lumped elements.
52
Transmission Line Fundamentals
have an impedance of 75Ω per meter. This is different from the resistance, where
the transmission line length does matter. For example, a trace might have a resistance of 0.1 Ω/inch.
The delay (tpd) has units of time (typically picoseconds or nanoseconds per
unit of length. A 10-inch-long trace having a delay of 180 ps per inch has a total
delay of 1.8 ns.
In Section 6.8 and the Problems we explore why the impedance does not depend on the lines length but the time delay does.
Equations (6.1) and (6.2) are only true when the effect of the resistance (R) is
very small compared to the effect of the inductance (L), and when the capacitor
losses (G) are very small compared to the effects of the capacitance (C). In fact,
these equations calculate the lossless impedance and lossless delay time because the
losses in the conductor (represented by R) and in the capacitor (represented by G)
are assumed to be small enough to drop out of the general equations for impedance
and delay. This is a good assumption for circuit board transmission lines carrying
signals with rise times of about 10 nS or faster [7].
For instance, the effect of the inductance is some 25 times higher than the effect of the series resistor for a pulse having a 1-ns rise time traveling down a 50Ω,
5-mil (127 μm)-wide half-ounce stripline trace on FR4. In contrast, the effect of the
inductance is only about eight times larger than the resistance when the rise time
increases to 10 ns.
6.4.1 Units and Electrical Length
The units of length associated with the values for L and C are important when using
these equations. For instance, field solvers usually give these values per unit length
(sometimes referred to as PUL), such as picofarads per centimeter or nanohenry per
inch. Total values for the entire length of the line are also sometimes given.
As an example, a 50Ω transmission line might have a PUL capacitance of 3 pF
per inch and an inductance of 6.5 nH/inch. A 10-inch-long line would have a total
capacitance and inductance equal to the PUL value times the total length: 30 pF
and 65 nH.
The electrical length of a transmission line having PUL values of 3.5 pF and 6.4
nH is 150 ps. It increases to 1.5 ns when the PUL values are 10 times as large. We
will see the advantage of looking at transmission lines in this way later on in this
chapter and in Chapters 11 and 12 when we discuss reflections and terminations.
6.5 How Does a Signal Travel Down the Line?
Because we took results from circuit theory rather than field theory, we have not
seen that energy travels as waves along transmission lines. We can develop an intuitive understanding of this by thinking about the transmission line circuit shown in
Figure 6.4 in the following way [8].
The circuit shows the resistance, inductance, and capacitance of a microstrip
over a return path such as a ground plane. For simplicity, the capacitor losses (the
G element in Figure 6.3) are not included.
6.5 How Does a Signal Travel Down the Line?
53
tpd
R1
tpd
L1
R2
tpd
L2
R3
L3
S
C1
C2
C3
Vs
V c1
V c2
Figure 6.4 Applying voltage to transmission line causes inductors and capacitors to charge in a
wave-like manner.
We will start off by first assuming that the switch is off, that there is no current
flowing in the circuit, and that the capacitors have all been discharged. We will also
assume that the line is infinitely long and so there are no reflections.
Closing the switch causes current to flow through L1 and R1, charging capacitor C1. However, the inductor and resistor prevent the capacitor from charging up
instantaneously. There is a time delay equal to tpd between the switch closing and
C1 being charged.
Once the delay is over, C1 is fully charged and current can now pass through
L2 and R2 to charge C2. After a delay of tpd seconds, C2 is fully charged and then
C3 starts to charge. This process continues until all of the capacitors along the line
have been charged.
By measuring the voltage across each capacitor, it appears as if a voltage wave
is traveling from left to right in the figure. By monitoring the current through the
inductors, it appears that a current wave travels along with the voltage wave.
This description of line behavior is simplistic. Rather than waiting for C1 to
become fully charged, current begins to flow in L2 while C1 is still charging, and
C2 can start to charge C3 before C2 is fully charged, and so on. This means that
the waveform gradually works its way down the line rather than doing so in steps
as we have described. However, if the segments are made small enough, the steps
become very small, and the observed behavior appears as a smooth wave traveling
down the line.
54
Transmission Line Fundamentals
6.5.1 What Voltage Gets Launched Down a Transmission Line?
Provided the line has low losses (which means that R and G in Figure 6.3 can be
ignored), the transmission line impedance appears as a simple resistance. The voltage and current waves at the input of an infinite length of line obey Ohm’s law, and
will have a ratio equal to Zo:
Zo =
Vi
Ii
(6.3)
where the launched voltage is called Vi, the incident voltage, and the launched current is called Ii, the incident current.
Equation (6.3) cannot be used when the impedance changes along the line or
at its end, which gives rise to signal reflections. Signal reflections are waves that
travel along the line independently of the launched wave. These waves combine in
complex ways, but for lines with no reflections, (6.3) shows that the characteristic
impedance determines the relationship between the voltage and current along the
line.
Because the impedance of a lossless transmission line acts like a simple resistance, the resistance voltage divider principle can be used to find the launched voltage if the power supply voltage (Vcc) is known. This is illustrated in Figure 6.5 and
in (6.4), where the impedance of the driver (such as an ASIC I/O driver) is Zg, and
the transmission line characteristic impedance is Zo.
Vi = Vcc
Zo
Zo + Zg
(6.4)
This model shows how the power supply voltage, line impedance, and I/O
driver impedance interact:
•
For a given driver impedance and Vcc value, a higher voltage will be launched
if the transmission line characteristic impedance increases. This can be used
to improve the received noise margins by increasing the voltage at the receiver, but care must be taken to prevent excessive reflections and overshoots.
Vi =
Zg = 50Ω
Zo
× Vcc = 1.65V
Z o + Zg
Zo = 50Ω
V cc= 3.3V
Figure 6.5 Voltage divider between the generator and transmission line impedances determines
the launched voltage (Vi). The line is long enough that the load is not a factor.
6.6 How Does the Current in the Return Path Behave?
55
•
The actual value of Vcc is important in determining the launched voltage.
Sags on Vcc caused by many drivers switching all at once (simultaneous
switching noise, described in Chapter 1) result in lower launched voltages.
This can lead to reduced receiver noise margins, and because the amount of
sag depends on the number of lines switching, the reduction in noise margin
becomes data dependent.
•
A higher voltage will be launched when low output impedance output drivers
are used. This may initially increase the received voltage, but care must be
taken because drivers with too low of an impedance will cause unacceptably
high reflection and overshoot voltages.
6.6 How Does the Current in the Return Path Behave?
We have seen how current flows down the transmission lines’ signal conductor, but
we have not seen how the current flows in the return.
We might think that the signal must first reach the load before current begins to
flow in the return, but in actuality current starts flowing in the return path as soon
as the signal is launched down the signal line.
We can see this by examining the microstrip shown Figure 6.6 and following a
current pulse (i) as it travels down the signal line. The current pulse reaches the first
capacitor, causing it to charge. To complete the circuit, the capacitor’s displacement current flows in the return path, back to the source. This occurs immediately,
before the signal has reached the load. Displacement current flows again when the
signal current reaches the next capacitor, and the process continues as the pulse
makes its way down the line. By monitoring the return path we observe that return
current continues to flow as each successive capacitor is charged. Because the distance between each capacitor is so small, a steady, uninterrupted current flows in
the return path. As illustrated by the arrows, this current moves in lockstep with
the signal current, but because it is flowing in the opposite direction it appears to
be a mirror image of it.
6.6.1 How Does Current Flow in Return Paths That Are Not Ground?
In the previous example the return path was a single plane, tied to ground, but the
return could have been a power plane rather than ground. The analysis for that situation is not presented here, but is similar to the analysis presented next for striplines
when one of the return paths is ground and the other is a voltage plane.
i
Vcc
i
Figure 6.6 Current in the return path is coincident with current i flowing in the signal trace.
56
Transmission Line Fundamentals
The power plane can either be connected to the same supply powering the I/O
driver, or it could be an unrelated supply that powers the ASIC core or other logic
on the board. As we will see, the return paths are very different in these situations.
6.6.2 What Is the Behavior When One Return Path Is a Related Power Plane?
An ASIC driving a stripline is shown in Figure 6.7. The I/O driver drives the line to
a logic high by closing the switch and connecting the resistance (which represents
the on resistance of the driver) to the ASIC I/O power supply.
The stripline is perfectly centered between a ground plane and a power plane
that connects to the ASIC I/O driver. This makes the capacitance (and inductance
not shown) from the trace to the ground plane identical to the capacitance (and
inductance) from the trace to the power plane.
This is not necessarily the case in practice. For example, with offset stripline
the trace is closer to one plane than the other, increasing the coupling to that plane
and decreasing it to the other plane. We can develop our explanation with that
configuration, but to simplify things, we will assume the capacitances (and the
inductances) to each plane are equal.
We start by assuming that the driver output has been at 0V long enough to fully
discharge all the capacitors between the trace and ground. Because the line is at 0V,
the capacitors formed with the power plane are all charged to the plane voltage.
From Figure 6.7 we see that once the driver switches to a logic high it launches
1.5V down the 50Ω transmission line. Solving for Ii with (6.3), we find that the
launched current is 30 mA.
Since the capacitances are equal, the same amount of return current flows in
the two planes. As expected, 15 mA flows in the ground plane back to the driver.
Figure 6.7 shows that the power plane current is actually the result of the driver
discharging the upper capacitors, which were charged when the line was held at
0V. The discharge path is through the ASIC power rail to the pull-up in the driver
and out to the line.
15 mA
15 mA
Power plane
30 mA
I/O
supply
15 mA
Ii =
= 30 mA
15 mA
V i = 1.5V
15 mA
50Ω
signal
trace
15 mA
Ground plane
Figure 6.7 Return path current when stripline is formed between a power plane connected to the
I/O driver and a ground plane.
6.6 How Does the Current in the Return Path Behave?
57
This analysis suggests that to properly manage the return currents it is not
necessary to decouple the I/O supply. In fact, to achieve the proper trace impedance, enough capacitance must be placed between the power and ground planes so
that the pair has an impedance very much less than the parallel impedance of all
the traces sharing the return. For instance, across frequency the decoupling should
have an impedance well under 0.5Ω if ten 50Ω traces were routed as in Figure 6.6.
This is explored in the Problems.
6.6.3 How Does the Return Current Flow When the Power Supply Is Not the
Signaling Voltage?
In Figure 6.8 we see a 50Ω signal trace sandwiched between a ground plane and a
power plane. In this situation the power plane does not connect to the pins on the
ASIC powering the I/O drivers. Instead, the stripline is formed with a plane carrying
an “unrelated voltage,” such as the power supply for the ASIC’s internal logic. For
instance, the internal ASIC logic supply might be 1.8V and the I/O supply is 3.3V.
Forming the stripline with the 1.8-V internal logic plane causes the capacitors (and the inductors, not shown) to be created between that power plane and
ground. Once again, we will assume that the trace is perfectly centered between the
two planes so that the capacitance to each is identical.
Figure 6.8 illustrates that, as in the previous case, the launched current evenly
divides between the two return planes. However, in this case, to discharge the upper capacitors, the 15 mA flowing in the power plane must flow through both
power systems before it can make its way back to the driver.
In general, the power supplies are resistive, and for frequencies above a few
tens of hertz, they are also inductive. The inductance prevents them from being an
effective path for high-frequency return current. In practice, the current mostly returns by way of the circuit board’s interlayer and decoupling capacitors. Interlayer
capacitance is the parallel plate capacitance naturally formed between a power and
ground plane. Each of the planes forms one of the capacitor plates, and the laminate is the dielectric.
15 mA
15 mA
Unrelated
supply
Power plane
30 mA
I/O
supply
Ii = 30 mA
15 mA
V i = 1.5V
15 mA
50Ω
signal
trace
30 mA
15 mA
15 mA
Ground plane
Figure 6.8 Return path current when stripline is formed when using a power plane not connected
to the I/O supply. Adequate decoupling capacitance is necessary to properly steer the return current.
58
Transmission Line Fundamentals
Generally, the path through the interlayer capacitance is low inductance, but
the capacitance is small (often in the tens of nanofarads range). This means that in
most cases the signal return currents flow through the decoupling capacitors explicitly placed between the power and ground planes. Unfortunately, the path through
the decoupling capacitors is inductive, which raises their impedance and makes
them increasingly less effective at high frequency. This is an important reason why
using an unrelated power supply to form a transmission line is poor practice and
should be avoided, especially when working with fast rise time signals since these
signals have high frequency harmonics that make achieving a low impedance AC
short difficult. As illustrated in the Problems, this means that the decoupling requirements for Figure 6.8 are much more stringent than those of Figure 6.7.
6.7 When Does a Conductor, Via, or Connector Pin Act Like a
Transmission Line?
As a rule of thumb, interconnect will act as a transmission line when the electrical
length of the line [as given in (6.2)] is longer than half the signal rise time [9–11].
Here we refer to the signal rise time to mean either the rise or the fall time, whichever is the fastest. The relationship is shown in:
LC ≥ 0.5tr
(6.5)
For instance, a line having a delay of 150 ps would be considered electrically
long (and so would exhibit transmission line behavior) when carrying signals having rise times of 300 pS or less.
More conservative estimates are limits of 0.2 [6] to 0.125 [12] rather than 0.5.
From (6.5) we see that:
•
It is the signal’s rise time relative to the interconnect length that determines if
a line is lumped or acts as a transmission line.
•
The signal pulse width and duty cycle are not factors in determining if a line
exhibits transmission line behavior such as reflections. Slow repetition pulses
with sharp edges are just as likely to trigger transmission line effects as are
narrow pulses with fast edges. The pulse width is important in determining
what the signal looks like as it travels up and down the line, but it does not
determine if reflections will occur.
A common signal integrity mistake is to assume that test logic or other lowspeed diagnostic interconnect used during debug or manufacturing test and configuration need not be analyzed for signal integrity. Unfortunately, although these
signals may be the slowest in the system, they are often generated by the same
circuit technology that creates the high-performance signals. The result is a sharp
rise time pulse with a slow repetition rate. Unless managed, this creates signal reflections and can lead to false triggering.
As explored in the Problems, this rule of thumb can also be used to decide if an
interconnect should be modeled as a lumped RLC circuit or as a transmission line.
6.8
Main Points
59
6.7.1 Why Does the Signal Rise Time Matter But Not the Pulse Width?
We have seen how the relationship between the length of line and the signal rise
time determines if the interconnect acts as a transmission line, but we have not explained why the pulse width is not a factor.
Assuming that the transmission line resistance R is very small, circuit theory
shows that the voltage drop across the inductor is determined by the amount of
time required by the current to change. This is illustrated in:
Vl = L ×
change in current
di
= L×
change in time
dt
(6.6)
For a given change in current (di), the voltage drop Vl is affected by the total
inductance L and the change in time (dt). As the voltage drop across the inductors
becomes smaller, the individual capacitors in Figure 6.4 increasingly act as if they
are connected in parallel. In this situation the line is no longer a distributed RLC
circuit, but is a single lumped RLC (or, if the voltage drop across the inductor and
resistor is very small, simply a lumped C circuit).
However, the capacitors become separated from one another by the inductor’s
voltage drop when it is high enough to no longer be negligible. In this situation the
inductors and capacitors are distributed along the line, and the line exhibits distributed RLC (transmission line) behavior.
This means that, because a short piece of interconnect does not have much inductance, the signal must have a very fast rise time before the voltage across the inductors becomes noticeable and the line acts like a transmission line. On the other
hand, if the interconnect is long, the inductance will be high, so even a slow rise
time signal can cause a noticeable voltage drop across the total inductance, causing
transmission line behavior. This matches our intuition. A short trace will not act
like a transmission line unless the signal rise time is very fast, but a long trace will
act like one even to slow rise time signals.
6.7.2 Why Does Propagation Delay Time Depend on the Line Length, But
Impedance Does Not?
It may seem natural to assume that the impedance and the propagation delay time
should depend on the line’s length, but, in fact, the impedance does not depend on
the length of line, while the propagation delay does. This comes about because (as
illustrated in the Problems) in (6.1) the impedance is the ratio of the inductance to
capacitance, while the propagation delay in (6.2) is the product of them. Because of
this, the length units cancel when calculating the impedance, but they do not in the
propagation delay calculations.
6.8
Main Points
•
A transmission line is comprised of two or more conductors: the signal and
its return.
60
Transmission Line Fundamentals
•
The return path determines the signal trace capacitance, inductance, and
resistance.
•
A transmission line can be modeled as a chain of very small RLCG elements.
•
The transmission line circuit elements are usually given as per unit length.
•
The transmission line capacitance and inductance cause the line to possess
impedance and impose a delay.
•
The transmission line impedance is independent of length, but the length
determines the line’s propagation delay time.
•
Current flows in the return path simultaneously with the signal current as a
mirror image.
•
The signal rise time determines if a line will behave as a transmission line or
as a lumped circuit.
•
A line will act as a transmission line if its electrical length is greater than
half the signal rise time. In some circumstances even shorter lines will act as
transmission lines.
•
Voltage divider action between the driver impedance and the transmission
line impedance determines the value of the launched signal voltage.
Problems
Answers to these problems are available on the Artech House Web site at http://
www.artechhouse.com/static/reslib/thierauf/thierauf1.html.
6.1
6.2
6.3
6.4
6.5
6.6
An ASIC drives a 50Ω transmission line with a driver having an output
impedance of 30Ω. The driver is connected to a 2.2-V supply. What current
is drawn from the supply when the driver switches from a logic low to a
logic high?
In the above setup, what voltage gets launched?
An ASIC has 30Ω, 50Ω, and 75Ω drivers in its library, and they can be
connected to either a 3.3-V or a 2.2-V supply. Which driver impedance and
supply voltage should be specified if the goal is to launch no less than 1.2V
but no more than 1.7V down a 50Ω transmission line?
An ASIC drives a 32-bit bus from a 3.3-V supply. The driver impedance can
be set to either 50Ω or 75Ω, and the circuit board traces can be created with
either of these impedances. When the drivers switch, it is necessary to limit
the total current drawn from the I/O power supply to under 750 mA. Using
only the ASIC total I/O current as the design consideration, what driver and
transmission line impedance should be chosen?
Using the results from Problem 6.3, which combination of driver and power
supply creates the least amount of on-die heat?
Find the propagation delay and impedance for a half-meter-long transmission line having L = 280 nH/meter and C = 118 pF/meter.
Problems
61
6.7
A circuit board trace has a total capacitance of 20 pF and a total inductance
of 50 nH. What is the propagation delay for this interconnect, and what is
the fastest rise time that can be sent down this trace before transmission line
effects occur?
6.8 Show how the propagation time but not the impedance scales with distance
by computing the delay and impedance for a 1-meter and a 5-cm-long transmission line. Assume that L is 280 nH/m and C is 110 pF/m.
6.9 A 50Ω transmission line has a propagation delay of 174 ps. What are the
line’s capacitance and inductance?
6.10 The trace routed to a termination device creates a half-inch stub. Assuming that the signal rise time is 350 ps and the line has a propagation delay
of 166 ps/inch, should the stub be modeled as a transmission line or as a
lumped capacitor?
6.11 Ten 50Ω stripline traces are routed as in Figure 6.6. Determine the minimum number of 10-nF decoupling capacitors that must be placed between
the power and ground planes when the signal has a 1-ns rise time. Assume
that when soldered to the circuit board each capacitor has an equivalent
series inductance of 700 pH.
6.12 Apply the results from Problem 6.11 to the configuration shown in Figure
6.8 and determine the number of needed capacitors.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
Young, B., Digital Signal Integrity, Upper Saddle River, NJ: Prentice-Hall, 2001, p. 326.
Miner, G. F., Lines and Electromagnetic Fields for Engineers, New York: Oxford University
Press, 1996.
Ramo, S., et al., Fields and Waves in Communication Electronics, 3rd ed., New York: John
Wiley & Sons, 1994.
True, K. M., “Data Transmission Lines and Their Characteristics,” National Semiconductor Application Note 806, February 1996.
Matick, R. E., Transmission Lines for Digital and Communication Networks, New York:
IEEE Press, 1969, p. 7.
Baloglu, H. B., Circuits, Interconnections, and Packaging for VLSI, Reading, MA: Addison-Wesley, 1990.
Johnson, H. W., and M. Graham, High-Speed Digital Design, Upper Saddle River, NJ:
Prentice-Hall, 1993.
Skilling, H. H., Transient Electric Currents, New York: McGraw-Hill, 1952, p. 262.
Blood, W. R., MECL System Design Handbook, Motorola Semiconductor Products, 1988.
Buchanan, J. E., Signal and Power Integrity in Digital Systems, New York: McGraw-Hill,
1996.
Hart, B. L., Digital Signal Transmission Line Circuit Technology, Berkshire, U.K.: Van
Nostrand Reinhold, 1988.
Royle, D., “Rules Tell Whether Interconnections Act Like Transmission Lines,” EDN, June
23, 1988, pp. 131–136.
CHAPTER 7
Understanding Microstrip and Stripline
Transmission Lines
7.1
Introduction
We saw in Chapter 5 how circuit board traces are arraigned as either microstrip or
stripline transmission lines. Figure 7.1 shows that microstrips are suspended over
a single plane at a height h, while striplines are sandwiched between two reference
planes that are separated by a distance b. From Chapter 6 we know that these
planes act as signal returns and can either be ground or power planes. As shown
Figure 7.1, the microstrips are usually covered with a solder mask, and the circuit
board may be hermetically sealed with an encapsulate (not shown).
By thinking of transmission lines as circuit elements rather than explaining
them with field theory, previous chapters did not show how the physical differences
in the construction of stripline and microstrips affect their electrical characteristics.
These factors are addressed in this chapter.
7.2
DC Resistance and Resistivity
When measured with an ohm-meter, the resistance of a circuit board trace (the
trace’s DC resistance) will usually be a few ohms or less, but it climbs to much higher
values when measured at high frequency. As is discussed in Chapter 8, this resistance increase means that conductor losses are much higher than at low frequencies.
The trace DC resistance depends on the trace resistivity (with units of ohmmeter), cross-sectional area (thickness and width), length, and temperature.
In Chapter 5 Table 5.5 showed that the resistivity of copper depends on how it
is formed. A typical value is 1.75 × 10−8 Ω-meters at room temperature.
The trace resistance increases with temperature. Although the exact amount
depends on the copper alloy, typically the resistance increases by just over 4% for
every 10°C rise in temperature [1]. This is a greater than 25% increase in DC resistance when the temperature increases from 25°C to 85°C. Because larger resistance
increases conductor losses, the signal integrity engineer should account for the effects of temperature when using field solvers to create worst-case trace models.
63
64
Understanding Microstrip and Stripline Transmission Lines
w
b
t
h
Microstrip
Stripline
Figure 7.1 Microstrip and stripline traces of width w, thickness t, and distance h or b.
Field solvers sometimes use conductivity (the reciprocal of resistivity, with units
of Siemens-meter) rather than resistivity to specify the metal characteristics. Depending on the alloy, at room temperature the conductivity of copper is about 5.7
× 107 S-m.
Figure 7.2 shows the room temperature resistance for half-ounce and 1-ounce
perfectly rectangular copper traces. The resistance increases dramatically for halfounce traces under 10 mils and 1-ounce traces less than 15 mils. The resistance is
higher than shown if the trace is trapezoidal. Equations for calculating trace resistance appear in Chapter 17.
7.3 Stripline and Microstrip Capacitance
As illustrated in Figure 7.3, a parallel plate capacitor is formed between the microstrip trace (as the upper plate of the capacitor) and the ground plane (as the
Trace Width (mm)
0.00
0.13
0.25
0.38
0.51
0.63
0.35
0.14
Half-ounce
0.12
0.25
0.10
0.20
0.08
0.15
0.06
Ohms/cm
Ohms/inch
0.30
One-ounce
0.10
0.04
0.05
0.02
0.00
0.00
0
5
10
15
20
25
Trace Width (mils)
Figure 7.2 Room temperature (25°C) DC resistance of half-ounce (0.65-mil-thick) and 1-ounce
(1.2-mil-thick) rectangular copper traces.
7.3 Stripline and Microstrip Capacitance
65
εr
h
Figure 7.3 Trace over a ground plane creates a parallel plate capacitor.
bottom plate). In stripline the coupling appears to both planes, even if one is a
power rather than a ground plane.
The fringing capacitance from the trace edges along with the capacitance between the trace bottom and the ground plane is not shown.
Increasing the distance h from the trace to the ground plane or reducing the
dielectric constant (εr, but, as mentioned in Chapter 1, sometimes called Dk) decreases the capacitance.
A typical dielectric constant (and the way in which they change with frequency)
for three FR4 laminate systems is shown in Table 7.1.
Because the polarization of the molecules in the dielectric is affected by frequency and temperature [5], the dielectric constant generally falls as frequency
increases, but in some laminates the dielectric constant will increase again after
reaching a minimum value at moderately high frequencies. FR406 exhibits this
behavior: εr falls from 4.6 at 1 MHz to 4.28 at 500 MHz and then climbs slightly
to 4.29 at 1 GHz. The rate of the change depends on the resin composition and the
amount of glass (the glass-to-resin ratio, introduced in Chapter 5).
7.3.1 What Is the Effective Dielectric Constant?
Because they are formed in one dielectric (or two that are very similar), striplines
are said to have a homogeneous dielectric and have a dielectric constant equal to εr.
The solder mask covering a microstrip has a dielectric constant close to typical
values for FR4, so microstrips are often thought of has having only two distinct
dielectrics: the laminate/solder mask and air. Because these two dielectrics have
different dielectric constants, the microstrip dielectric is not homogeneous, but it
is possible to think of the blend as having one aggregate value. The microstrip effective dielectric constant (εr_eff) is an equivalent value that takes into account the
value of εr of the laminate and the microstrip width and thickness and its height
above the return path. These physical factors are important because they determine
Table 7.1 Dielectric Constant and (Loss Tangent) for Three
FR4 Type Laminate Systems for a Glass-to-Resin Ratio of 44%
Laminate
1 MHz
500 MHz
1 GHz
FR404
4.6 (0.025)
4.26 (0.014)
4.25 (0.014)
FR406
4.6 (0.023)
4.28 (0.014)
4.29 (0.014)
FR408
3.8 (0.010)
—
3.7 (0.010)
Source: [2–4].
66
Understanding Microstrip and Stripline Transmission Lines
how much of the signal energy travels in air and how much travels in the laminate
and how much each influences the capacitance and (as we will see in Section 7.8)
the signal delay.
The effective dielectric constant will have a value between 1 (for those fields
traveling only in the air) and the value of the laminate itself. This means that the
effective dielectric constant will always be lower than the dielectric constant of the
underlying laminate, so signals propagating along microstrip experience a lower
dielectric constant than striplines. This is true even when they are fabricated with
the same laminate materials. As we will see, this makes the electrical characteristics
of microstrips quite different from striplines.
7.3.2 What Are the Differences Between Microstrip and Stripline
Capacitances?
Figure 7.4 plots the capacitance for several microstrip and stripline traces for three
impedances and traces of various widths. The traces are perfectly rectangular, εr is
4.2, and the microstrips are covered with a 1-mil-thick solder mask that also has an
εr of 4.2. Chapter 17 shows how to calculate capacitance using trace dimensions or
trace impedance.
The graph shows that high impedance traces have less capacitance than low
impedance traces, and for a given impedance, the capacitance of stripline traces is
higher than microstrip traces.
Width (mm)
Capacitance (pF/inch)
4.25
0.10
0.15
0.21
0.26
0.31
0.36
0.41
0.46
0.51
0.56
1.80
1.70
40Ω
4.00
1.60
3.75
1.50
3.50
1.40
50Ω
3.25
1.30
3.00
1.20
60Ω
2.75
Capacitance (pF/mm)
0.05
4.50
1.10
1.00
2.50
2
4
6
8
10
12
14
16
18
20
22
Width (mils)
Figure 7.4 Capacitance of rectangular solder mask microstrip (plotted with dotted lines) and stripline (solid lines) traces of various widths. Raw data from the Polar Si9000 2D field solver [6].
7.3 Stripline and Microstrip Capacitance
67
For instance, the capacitance of a 5-mil (0.13-mm)-wide, 60Ω microstrip is
about 2.6 pF/inch, but it increases to about 4 pF/inch for a 40Ω microstrip of the
same width. A 40Ω stripline of any width has a capacitance of about 4.3 pF/inch,
which is about 10% higher than the microstrip capacitance.
7.3.3 How Does Frequency Alter the Loss Tangent?
In general, the loss tangent (introduced in Chapter 1) will have its highest value at
low frequencies, and the value of the loss tangent falls as frequency increases. This
trend is evident for three FR4 laminate systems in Table 7.1. For instance, the loss
tangent of the first laminate is 0.025 at 1 MHz, but falls to 0.014 for higher frequencies. As in this case, once reaching a high enough frequency, the loss tangent
will often remain nearly constant and in some cases it may begin to climb again.
Moisture uptake (described in Chapter 8) can cause the loss tangent to increase in
high-humidity environments.
7.3.4 What Is Conductance and How Is It Determined?
Conductance is used in many SPICE lossy transmission line models to account for
stripline and microstrip dielectric losses, and this chapter’s Problems show how
to create SPICE lossy transmission line models. In Chapter 8 we will see how to
directly determine loss from the loss tangent without using the conductance. However, because conductance is so widely used in simulators, it is worthwhile to understand how its value is determined.
From Chapter 6 we recall that the conductance (G, units of Siemens/length)
is modeled as a resistor in parallel with the trace capacitance. Higher values of
conductance represent higher dielectric losses. As shown in (7.1), the conductance
value is determined by the trace capacitance (C), the frequency (f ), and the value
of the loss tangent (LT).
G = 6.28 × C × f × LT
(7.1)
Evidently traces with larger capacitance and loss tangent values have correspondingly larger values for G (and so, higher dielectric losses). This means that
dielectric loss is highest for low impedance traces (because these have larger capacitance), and the loss increases with higher loss tangent values.
As we just saw, for a given impedance the capacitance of a stripline is the same
for all trace widths, but it changes for microstrip widths. This means that for a
given impedance, the dielectric loss is the same for striplines regardless of its width,
but the dielectric loss will be different for various widths of microstrip. This assumes the loss tangent has the same value everywhere on the circuit board.
The effects with microstrips are small, especially if the microstrip is covered
with solder mask, and although not evident from the equation, because the loss
tangent for air is lower than that of the laminate, for any given impedance G is
lower for microstrip than for stripline.
For instance, on FR4 (εr = 4.2, LT = 0.02), G is just over 17m S/m for a 50Ω,
half-ounce stripline of any width at 1 GHz. At that same frequency it is slightly
more than 15m S/m for a 5-mil-wide microstrip that is covered by 1 mil of solder
68
Understanding Microstrip and Stripline Transmission Lines
mask. It falls to about 14.75m S/m when the 50Ω solder mask–covered microstrip
is 10 mils wide. It is even less (just over 9m S/m) when the 10-mil microstrip has
an impedance of 75Ω.
This example and (7.1) show the advantage of using wide, high-impedance microstrips to route high-speed signals, especially when an inexpensive circuit board
having a high loss tangent is used. We will return to this when discussing dielectric
losses in Chapter 8.
7.4 Stripline and Microstrip Inductance
Circuit board traces exhibit parallel plate inductance with the ground plane. The
inductance is determined by the height and width of the trace: Increasing the height
(h and b in Figure 7.1) causes the inductance to become larger, while increasing the
trace width (w) lowers the inductance. The skin effect causes the inductance to fall
slightly at high frequency as the internal inductance is reduced. In fact, at very high
frequencies the inductance is mostly determined by the shape and dimensions of the
trace. These things are described in more detail in Section 7.5.
The high-frequency inductance for several microstrip and stripline traces is
presented in Figure 7.5. The traces are perfectly rectangular, the dielectric constant
is 4.2, and the microstrips are covered with a 1-mil-thick solder mask having a
dielectric constant of 4.2.
Inductance (nH/inch)
12
0.56
4.8
11
4.4
10
0.10
0.15
0.21
0.26
0.31
0.36
0.41
0.46
0.51
60Ω
4.0
9
3.6
50Ω
8
3.2
7
Inductance (nH/mm)
Width (mm)
0.05
2.8
40Ω
6
2.4
2
4
6
8
10
12
14
16
18
20
22
Width (mils)
Figure 7.5 Inductance at a very high frequency of rectangular traces of various widths. Striplines
are shown with solid lines; the dotted lines are microstrips. Raw data from the Polar Si9000 2D field
solver [6].
7.5 How Does the Skin Effect Change the Trace’s Resistance?
69
Low-impedance traces are seen to have lower inductance than high impedance
traces. For example, the 40Ω stripline inductance is about 7 nH/inch for any width
versus about 10.5 nH/inch for the 60Ω stripline.
Also evident is that for a given impedance and width, the stripline has a higher
inductance than microstrip. For instance, a 5-mil-wide 50Ω stripline has an inductance of 8.7 nH/inch, while the microstrip is 8 nH/inch.
7.5 How Does the Skin Effect Change the Trace’s Resistance?
At low frequencies the signal current flows through the entire cross-section of a
trace. However, at high frequencies the skin effect causes current to only flow near
the trace surface in a region defined by the skin depth. The skin depth (or the depth
of penetration) is large for low frequencies, which is why the current flows through
the entire trace cross section at DC.
However, at a high frequency the resistance of the trace and the return path (the
loop resistance, Rac) increases as the square root of the frequency. This is shown in:
Rac _ f 2 = Rac _ f 1
f2
f1
(7.2)
For instance, increasing the frequency by a factor of 4 from 500 MHz to 2 GHz
causes the resistance to increase by 2.
The skin effect becomes apparent in 1-ounce traces at roughly 14 MHz; it is
evident at about 65 MHz for half-ounce traces [7].
The loop resistance at 350 MHz (corresponding to the 3-dB bandwidth of a
1-ns pulse) is plotted for various widths of striplines and solder mask covered microstrips in Figure 7.6. For the laminate and the solder mask, the dielectric constant
is 4.2, and the loss tangent is 0.02. The solder mask is 1-mil thick.
The loop resistance for frequencies other than 350 MHz can be found by using
(7.2). For instance, Figure 7.6 shows the loop resistance of a 4-mil (0.11-mm)-wide
50 stripline to be about 0.8 Ω/inch. Increasing the frequency to 1.4 GHz causes the
resistance to increase by a factor of 2 (f1 = 350 MHz, f2 = 1.4 GHz, Rac_f1 = 0.8Ω),
to 1.6 Ω/inch.
7.5.1 What Is the Proximity Effect?
One property of the skin effect (the proximity effect) is that as frequency increases,
current in the return path gradually collects underneath the trace, rather than
spreading out and using the entire ground plane as it does at lower frequencies
[9–12]. For a given trace width, the spreading is greatest for traces further away
from the ground plane, which is why higher-impedance traces (which as we will see
in Section 7.7 are further away from the return plane) use more of the return path
metal than lower-impedance ones. This is shown in Figure 7.7.
70
Understanding Microstrip and Stripline Transmission Lines
Width (mm)
0.05
1.40
0.08
0.10
0.13
0.15
0.18
0.20
0.23
0.25
0.55
50Ω
0.50
1.27
1.14
0.45
60Ω
70Ω
0.40
50Ω
60Ω
70Ω
0.89
0.35
0.76
0.30
0.64
0.25
0.51
0.20
0.38
0.15
2
3
4
5
6
7
8
9
R ac (Ω/cm)
Rac (Ω/inch)
1.02
10
Width (mils)
Figure 7.6 Loop resistance as calculated by Linpar [8] at 350 MHz for perfectly rectangular 0.65-mil
(0.017-mm)-thick stripline (solid curve) and solder mask–covered microstrip (dotted curve).
100Ω
50Ω
Current extends out
beyond conductor edges
Figure 7.7 Cross-section of two microstrips showing the current in the ground plane gathering
underneath the traces at high frequency. The proximity effect cause more spreading for traces of the
same width that are a greater distance from the return.
Confining the return current to a small channel in the vicinity of the trace raises
the loop resistance. Conversely, the resistance will be lower if the current can spread
out along a larger section of the return plane (as it does for traces higher above
the plane). The net effect is that, for a given dielectric constant and trace width,
the loop resistance is lower for high-impedance traces at any given high frequency.
7.6 How Does the Skin Effect Change the Inductance?
71
7.6 How Does the Skin Effect Change the Inductance?
The total trace inductance of a trace or round wire is the sum of the inductance
internal to the trace (the internal inductance) and the inductance due to the trace
geometry (the external inductance) [5, 13].
The internal inductance is small and its value is reduced as the skin effect
causes the current to move away from the center of the trace and migrate outward,
toward the trace surface. This leaves fewer lines of magnetic flux present inside the
conductor, lowering the internal inductance. Because this behavior is driven by the
skin effect, the internal inductance decreases as the square root of the increase in
frequency (the same rate at which the resistance increases).
For instance, at 500 MHz the internal inductance of a 4-mil-wide 50Ω stripline
trace is about 5% of the total inductance. It falls to less than 1% at 20 GHz. Assuming that the capacitance does not change, this reduction in inductance causes
the trace impedance to be about 2% lower at 20 GHz than it is at 500 MHz.
7.7 Understanding Stripline and Microstrip Impedance
A trace’s dimensions (its width, thickness, and distance to the ground plane or
planes) determine its impedance. To understand how these interact, (7.3) (simplified
from [14]) calculates stripline impedance, and (7.4) calculates microstrip impedance. For clarity, (7.4a) is slightly modified from [15], and uses a more comprehensive equation for the effective dielectric constant (7.4b) [16].
More complex (and accurate) equations appear in Chapter 17, but these equations are simple enough to show how the trace dimensions interact and determine
the impedance. The traces are assumed to be perfectly rectangular, and the stripline
is assumed to be centered between the return planes.
Zo =
Zo =
εr _ eff =
⎛ b + w⎞
ln ⎜
⎟
εr ⎝ w + t ⎠
94
60
εr _ eff
⎛ 7.5h ⎞
ln ⎜
⎝ w + 1.25t ⎟⎠
εr + 1 εr − 1 ⎛
w ⎞
+
⎜
2
2 ⎝ w + 10h ⎟⎠
(7.3)
(7.4a)
(7.4b)
From these equations we see that:
•
When holding the distance to the trace (b or h) constant, the impedance
gradually goes down as the trace gets wider (w) or as it gets thicker (t).
72
Understanding Microstrip and Stripline Transmission Lines
•
•
When holding the width (w) and thickness (t) fixed, the impedance gradually
goes down when the distance to the ground plane (b or h) is made smaller.
The impedance goes up as the dielectric constant (εr for stripline, or εr_eff for
microstrip) goes down, and it does so as the square root of the change.
For instance, doubling the trace width from 5 mils to 10 mils but keeping the
spacing b constant at 13.5 mils causes the stripline impedance to fall by less than
half from 50Ω to 31Ω on FR4. In a similar way, reducing the dielectric constant in
half from 4.2 to 2.1 causes the stripline impedance to increase from 50 to nearly
71Ω (rather than doubling).
Because the units cancel, either inch-based or metric units may be used with
these equations to obtain the impedance. For impedances of 40Ω or higher, (7.3) is
accurate for striplines to better than 5%. Equation (7.4a) is accurate to within 10%
for microstrips without solder mask across that same impedance range when used
with (7.4b) to calculate εr_eff. Chapter 17 shows how to calculate the impedance of
solder mask–covered microstrip.
This chapter’s Problems further illustrate uses for these equations, but they are
plotted for FR4 (εr = 4.2) in Figure 7.8.
For stripline the trace bottom is assumed to be centered between the planes
at the distance shown. For the microstrip traces the distance is from the ground
plane to the trace bottom. The solder mask covers the microstrip by 1 mil and has
a dielectric constant of 4.2.
Figure 7.8 is for illustration only because manufacturing details such as the
actual shape of the trace and the composition of the laminate are not accounted for.
0.10
0.13
Trace width (mm)
0.15 0.18 0.20 0.23
34
32
30
28
26
24
22
20
18
16
14
12
10
8
6
4
2
0
70Ω
0.25
60Ω
50Ω
40Ω
70Ω
60Ω
50Ω
40Ω
3
4
5
6
7
8
9
0.85
0.80
0.75
0.70
0.65
0.60
0.55
0.50
0.45
0.40
0.35
0.30
0.25
0.20
0.15
0.10
0.05
0.00
Distance (mils)
Distance (mm)
0.08
10
Trace width (mils)
Figure 7.8 The distance (b or h) and trace width (w) necessary to obtain various impedances for
stripline (solid curve) and solder mask–covered microstrip (dotted curve). Data from the Polar Si9000
2D field solver.
7.8 Understanding Stripline and Microstrip Propagation Delay Time
73
7.8 Understanding Stripline and Microstrip Propagation Delay Time
In air, electromagnetic energy requires about 3.3 ns to travel the distance of 1 meter
(about 85 ps per inch), but as shown in (7.5), it takes longer to travel through a dielectric such as a circuit board laminate. The physics behind this is widely available
(for just one example, see [5]), but we will note here that as shown in Table 7.1, εr
for circuit board laminates is higher than the value for air (which is about 1).
tpd = 84.7 εr ps/inch
(7.5a)
tpd = 33.5 εr ps/cm
(7.5b)
In (7.5) tpd is the propagation delay time [in picoseconds/inch for (7.5a), and
picoseconds/centimeter for (7.5b)], and εr is the dielectric constant of the laminate
for stripline or the effective dielectric constant (εr_eff) for microstrip.
Because only one dielectric is involved with stripline (the dielectric is homogeneous), (7.5) shows that the delay is not affected by the trace dimensions or impedance. For example, the delay of a 4-mil-wide, 25Ω stripline will be identical to the
delay of a 10-mil-wide, 100Ω stripline formed on the same layer (where we assume
εr is identical for both traces). In fact, the delay for stripline traces is independent
of the trace geometry.
However, because the two dielectrics (air and the laminate) have such different
dielectric constants, this is not so for microstrips. In this case, the dielectric is not
homogeneous, and the trace width and impedance do influence microstrip delay.
To understand this, we notice from Figure 7.8 that obtaining a given impedance while making w larger requires the trace to be further away from the ground
plane. This means that more of the field lines travel in the air as compared to a
trace that is closer to the laminate. Since signals travel faster in air than in a laminate, this results in the delay being smaller. In fact, εr_eff becomes closer to 1 as the
trace is moved further away from the ground plane (that is, as h in Figure 7.1 is
made larger). From (7.5), lower values of εr_eff cause tpd to be less. Microstrip is
said to be faster than stripline for this reason.
Similarly, as is evident in Figure 7.8, obtaining a particular impedance for a
given trace width requires that h change. This means that, unlike stripline, the delay of a 25Ω microstrip will not be the same as the delay of a 100Ω microstrip. In
fact, the 25Ω microstrip will be closer to the laminate than the 100Ω microstrip,
making its delay longer.
7.8.1 How Do the Dielectric and Effective Dielectric Constants Affect Delay
Time?
Equations such as (7.4b) do not clearly show how the delay changes with trace
width and impedance. The delay for various trace widths is plotted in Figure 7.9 to
show this relationship.
Signal delay on microstrip is reduced slightly with increases in the trace width
and impedance, and microstrips are faster than stripline. For instance, on FR4
Understanding Microstrip and Stripline Transmission Lines
0.05
180
0.10
0.15
Delay (ps/inch)
175
Width (mm)
0.20
0.25
0.30
0.36
0.41
71
69
Stripline
170
67
165
65
Microstrip
160
63
Delay (ps/cm)
74
50Ω 61
60Ω
155
150
2
4
6
8
10
Width (mils)
12
14
59
16
Figure 7.9 Approximate delay of half-ounce-thick stripline and microstrip on FR4 (εr = 4.2). The
stripline of all widths and impedances has a delay of 174 ps/inch (68.5 ps/cm), but the microstrip
trace width and impedance factor into the microstrip delay. The solder mask thickness is 1 mil (25.4
μm).
the 3-mil-wide 50Ω microstrip is about 8% faster than any of the shown stripline
traces, and is about 12% faster for a 15-mil-wide trace.
Wide, high-impedance microstrips have smaller delay because they are further
away from the ground plane and more of the field lines travel in air. However, the
effect is small. Changing the impedance by 20% from 50Ω to 60Ω reduces the
delay by just over 1%.
The data shown is for a 1-mil (25.4-μm)-thick solder mask, which has a dielectric constant of 4.2. A thicker coating of solder mask causes the trace to be slower
(doubling the solder mask thickness from 1 to 2 mils causes the propagation time
to increase by about 3% on FR4). Increasing the microstrip thickness to 1-ounce
copper increases the delay by about 1%.
7.8.2 How Sensitive Is Delay Time to Changes in the Dielectric Constant?
The delay for a 3-mil (0.08-mm)-wide and 10-mil (0.25-mm)-wide 50Ω microstrip
and a stripline is shown in Figure 7.10 plotted against different values for the circuit
board dielectric constant (εr). The traces are perfectly rectangular half-ounce copper
with a 1-mil (25.4-μm) solder mask, whose εr is 4.2.
The delay difference between stripline and microstrip traces grows as the dielectric constant gets larger. Large dielectric constants also reduce the difference in
time of flight between wide and narrow microstrip traces.
7.9
Main Points
•
The trace and return path form parallel plate capacitors and inductors.
Main Points
75
210
84
All
widths
200
80
190
76
3-mils
wide
180
72
10-mils
wide
170
68
160
64
150
60
140
56
3.0
3.5
4.0
4.5
5.0
5.5
Delay (ps/ch)
Delay (ps/inch)
7.9
6.0
Dielectric Constant (εr )
Figure 7.10 Approximate delay for stripline (solid curve) and solder mask covered microstrips (dotted curve) fabricated on circuit boards of various dielectric constants. Stripline delay is not affected
by trace width, but microstrip delay is.
•
The distance between the trace and the return path determines the trace resistance, inductance, capacitance, and conductance.
•
The dielectric systems are not the same for microstrip and stripline traces,
making them have different electrical characteristics.
•
Transmission line loss is caused by trace resistance and leakage in the
dielectric.
•
The skin effect causes the AC loop resistance to increase and becomes apparent for half-ounce traces at about 65 MHz (lower for thicker traces).
•
For a given dielectric constant, the trace thickness, width, and distance from
the return path determine the trace impedance (Zo).
•
In striplines the trace propagation delay time (tpd) is only determined by the
dielectric constant, giving all striplines on a routing layer the same delay per
length.
•
Trace geometry factors in the microstrip delay time.
•
For striplines of a given impedance, the trace capacitance (C), inductance (L),
and conductance (G) are the same for all trace widths.
•
For microstrips of a given impedance, the trace capacitance, inductance, and
conductance are affected by the trace width, especially when the dielectric
constant is low or when the solder mask is thin or has a low Dk.
•
The skin effect causes a traces inductance to decrease as frequency increases.
76
Understanding Microstrip and Stripline Transmission Lines
Problems
Answers to these problems are available on the Artech House Web site at http://
www.artechhouse.com/static/reslib/thierauf/thierauf1.html.
7.1
What is the approximate impedance of a half-ounce 5-mil-wide stripline on
FR4 when the plane separation is 16 mils?
7.2 What is the approximate impedance of a half-ounce 8-mil-wide solder mask
microstrip on FR4 when the separation from the trace to the return plane
is 5.2 mils?
7.3 How will the board thickness change when a half-ounce, 50Ω stripline on
FR4 is changed from 5 mils wide to 8 mils wide?
7.4 How long will it take a signal to travel down a 12-inch (30.5 cm) long, 75Ω,
5-mil (0.127-mm)-wide, half-ounce stripline fabricated on FR4 (Dk = 4.2)?
7.5 A certain circuit board cannot be made too thick, yet it is necessary to
obtain the maximum number or routing channels possible. What influence
does this requirement have on selecting the trace width and impedance?
7.6 A 50Ω stripline is fabricated on FR4 (εr = 4.2). What will be the impedance
and delay time if the laminate is changed to one having εr = 3.75?
7.7 A 10-mil-wide, 50Ω stripline runs parallel to and on the same layer as a
75Ω, 5-mil-wide stripline. Assuming that the laminate is FR4 and both traces are a half-ounce thick, which trace will delay the signal the least?
7.8 The circuit board manufacturer has decided to reduce the thickness of each
stripline layer in a multilayer circuit board from 15 mils (0.38 mm) to10
mils (0.25 mm) as a way to meet the overall board thickness goal. What effect will this have on the trace impedance and time of flight?
7.9 A stripline is known to be precisely 31.5 inches (80 cm) long. A signal takes
5.4 nS to travel the entire trace length. What is the dielectric constant of the
laminate?
7.10 From a 2D field solver, Rac is found to be 7 Ω/meter (0.18 Ω/inch) at 100
MHz. What is the resistance at 1 GHz?
7.11 What is the conductance at (a) 100 MHz and (b) at 1 GHz for a 5-mil
(0.13-mm)-wide half-ounce 50 stripline on FR4 (Dk = 4.2 and LT = 0.02)?
7.12 Build a W line model for a 5-mil-wide, half-ounce 50Ω stripline on FR4.
References
[1]
[2]
[3]
[4]
[5]
[6]
The Chemical Rubber Company, CRC Handbook of Chemistry and Physics (1986–1987),
67th ed., Cleveland, OH: CRC Press, 1987, pp. E-9.
Polyclad Laminates, Inc.,” PCL-FR-226/PCL-FRP-226,” data sheet PCL-226, dated 10/01,
Isola Group, Chandler AZ.
Isola Laminate Systems, Inc., “FR408 Product Data Sheet,” No. 5035/9/99, LaCrosse, WI,
1999.
Isola Group, “Isola IS620,” data sheet IS620 rev 07-08-08, Chandler, AZ.
Miner, G. F., Lines and Electromagnetic Fields for Engineers, New York: Oxford University
Press, 1996.
Polar Si9000 Field Solver, Polar Instruments, Inc., http://www.polarisinstruments.com.
Problems
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
77
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Djordjevic, A. R., et al., LINPAR for Windows, Norwood, MA: Artech House, 1999.
Bertin, C. L., “Transmission-Line Response Using Frequency Techniques,” IBM Journal of
Research and Development, January 1964, pp. 52–63.
Faraji-Dana, R., and Y. L. Chow, “The Current Distribution and AC Resistance of a Microstrip Structure,” IEEE Journal of Microwave Theory and Techniques, Vol. 38, No. 9,
September 1990, pp. 1268–1277.
Pucel, R. A., D. J. Masse, and C. P. Hartwig, “Losses in Microstrip,” IEEE Transactions
on Microwave Theory and Techniques, MTT-16, No. 6, June 1968, pp. 324–350.
Pucel, R. A., D. J. Masse, and C. P. Hartwig, “Correction to Losses in Microstrip,” IEEE
Transactions on Microwave Theory and Techniques, Vol. MTT-16, No. 12, December
1968, p. 1064.
Matick, R. E., Transmission Lines for Digital and Communication Networks, New York:
IEEE Press, 1969.
Bakoglu, H. B., Circuits, Interconnections, and Packaging for VLSI, Reading, MA: Addison-Wesley, 1990.
Kaupp, H. R., “Characteristics of Microstrip Transmission Lines,” IEEE Transactions on
Electronic Computers, Vol. EC-16, No. 2, April 1967, pp. 185–193.
Scheider, M. V., “Microstrip Lines for Microwave Integrated Circuits,” Bell System Technical Journal, May-June 1969, pp. 1421–1444.
CHAPTER 8
Signal Loss and the Effects of Circuit
Board Physical Factors
8.1
Introduction
In previous chapters we assumed that the copper traces were rectangular and had
smooth surfaces and that the stripline traces were embedded in a single block of
dielectric material (the dielectric was homogeneous).
These ideal assumptions are not valid in practice. Copper surfaces are rough,
often traces are trapezoidal in shape, and the glass-to-resin ratio (see Chapter 5)
determines the actual dielectric constant. All of these can alter the trace impedance,
delay, and loss characteristics, and it is important for the signal integrity engineer to
determine when it is necessary to account for them in simulation or when to discuss
them with circuit board fabrication shops.
8.2 What Are Transmission Line Loss and Attenuation?
Loss is the mechanism that reduces (attenuates) a signal and in circuit board traces
is caused by resistance in the conductor and leakage in the circuit board laminate.
Distortion is the change in pulse shape caused by the loss unequally affecting each
of the harmonics making up the signal.
8.2.1 What Effects Do Losses Have on Signals?
To understand loss, we must think about signals in the frequency domain. As described in Chapter 1, a pulse is made up of many frequency components (harmonics). Because the amount of loss depends on the frequency, each harmonic making
up a pulse will be attenuated differently as the pulse travels down the line. In fact,
each successively higher harmonic will be attenuated more than lower-frequency
ones. This means that when the harmonics recombine at the load their amplitudes
and phase are not in the same proportion as when they were launched. The pulse
has become distorted.
79
80
Signal Loss and the Effects of Circuit Board Physical Factors
As shown in Figure 8.1, the attenuation of the high-frequency components
cause the received pulse to become rounded and reduced in amplitude, and the
base of the pulse smears and becomes wider than the base of the transmitted pulse.
The progressive distortion of a 1-V, 400-ps pulse stream as it progresses down
a 1-meter (39-inch)-long transmission line is illustrated in Figure 8.2. The effect on
the pulse is evident when compared to the undistorted pulse. In fact, the pulse base
has become wide enough to collide with the rising edge of the succeeding pulse.
This effect (dispersion) can severely distort a pulse stream, which is of particular
concern in serial signaling. For instance, dispersion severe enough to cause a pair
of back-to-back pulses (such as a 0110 pattern) to interfere with one another may
not be great enough to cause significant interference when the pulses are widely
separated (for example, a 1001 pattern). This leads to “data-dependent” signaling
errors.
8.2.2 How Is Loss Specified?
In signal integrity work, loss is defined as the ratio of the received voltage to the
transmitted voltage. As shown in (8.1), the ratio is given in terms of the common
logarithm and has units of decibels (dB).
⎛ V
⎞
Gain in dB = 20 × log10 ⎜ received ⎟
⎝ Vtransmitted ⎠
(8.1)
When calculating loss, decibels are used rather than a simple percent because it
is only necessary to add together the losses in decibels of the various parts to find
the total loss. As we will see, microstrip and stripline losses come in two parts, and
the total loss in dB is simply the sum of the two.
Input Pulse
1.0V
1m (39 inches)
0.8V
0.6V
0.4V
0.2V
0.0V
0 ns
5 ns
10 ns
15 ns
20 ns
Figure 8.1 Effects of loss created by a 1-meter (39-inch)-long microstrip on a sharp edged, 5-nSwide pulse (dotted curve). The pulse exiting the line (solid curve) is smaller, rounded, and wider at
its base.
8.3 What Creates Transmission Line Loss and How Is It Characterized?
1.0V
L = 0.25m
L = 0.50m
L = 0.75m
81
L = 1.0m
0.5V
0.0V
2.0 ns
4.0 ns
6.0 ns
8.0 ns
Figure 8.2 Transmission line loss causes pulse distortion and spreading of its base. The position and
shape of the pulse without distortion are shown for reference.
Because the logarithm of a fraction has a negative value, the gain will be negative when Vreceived is less than Vtransmitted. For instance, the gain is −14 dB when
Vreceived is 0.2V and Vtransmitted is 1V. However, it is customary to drop the negative
sign and say that the circuit has a loss (or the signal has experienced an attenuation) of 14 dB.
This equation is further demonstrated in the Problems, and Chapter 17 shows
how to convert from decibels to the ratio of the received and transmitted voltage.
The equation cannot be used to represent the power lost in dB, but by using the
received and transmitted currents (Ireceived and Itransmitted), it can show the current
attenuation.
8.2.3 Converting Between Decibels and Percentage Loss
The relationship between the loss in percent and in decibels is shown graphically in
Figure 8.3. The negative value is implied and the attenuation is shown as a positive
number.
A 3-dB loss is shown to mean that the signal voltage is reduced by just over
29%, leaving nearly 71% of the original signal intact at the receiver. The nonlinear
nature of the decibel scale is evident: increasing the loss from 1 to 10 dB reduces the
received amplitude by 2.8 times (89% to 32%) rather than 10 times.
8.3 What Creates Transmission Line Loss and How Is It Characterized?
Signals lose energy by heating the conductor resistance (represented by R in Figure
8.4) and by leakage currents heating the dielectric (G). As we saw in Chapter 7,
both of these change with frequency. For a circuit board, the capacitor dielectric is
the laminate or prepreg, and the conductor is a trace and the ground plane(s).
82
Signal Loss and the Effects of Circuit Board Physical Factors
100
Received
90
Signal Voltage (%)
80
70
60
50
40
30
20
Lost
10
0
0
5
10
15
20
Transmission Line Loss (dB)
Figure 8.3 Amplitude (in percent) to decibel conversion. A 3-dB loss is seen to represent a signal
swing of 70.8% at the receiver.
Alpha (α) is used to indicate loss, with a subscript showing that it is either the
dielectric (αd) or the conductor (αc) part. As shown in (8.2), the total loss in dB is
the sum of the two and is represented by αt.
αt = αc + α d
(8.2)
As Figure 8.5 shows for FR4 (εr = 4.2, LT = 0.02), at low frequencies losses in
the conductor contribute more to the total loss than dielectric losses. However, the
two losses increase at different rates, so at some frequency losses in the dielectric
overtake and eventually contribute more to the total than the conductor loss. In
this example the crossover point occurs at about 1 GHz.
This type of analysis is used to decide if a laminate system having dielectric
constants and loss tangent values lower than typical FR4 should be considered. Using Figure 8.5 as a specific example, a laminate having very low loss tangent values
(which lowers the dielectric loss) is not warranted if most of the signal harmonics
are below roughly 1.5 GHz.
8.3.1 Loss Differences Between Microstrip and Stripline Traces
The difference in loss of microstrips and striplines is evident in Figure 8.6. By picking one frequency (the 3-dB bandwidth of a 1-ns rise time pulse), the graph clearly
shows how the two physical parameters most easily adjusted by the signal integrity
engineer (trace width and impedance) influence loss.
The loss for narrow traces is higher in microstrips than stripline, but the situation reverses for wider traces, where stripline losses are higher than microstrip
8.4 How Do Impedance and Loop Resistance Affect Conductor Loss?
R
83
L
G
C
Figure 8.4 Transmission line circuit model. Resistance in the conductor and losses in the dielectric
cause signal loss.
0.31
0.8
0.7
αt
0.5
αd
0.20
0.16
0.4
0.12
0.3
Loss (dB/cm)
0.24
0.6
Loss (dB/inch)
0.28
αc
0.2
0.08
0.1
0.04
0.0
0.00
1 GHz
2 GHz
3 GHz
4 GHz
5 GHz
Frequency
Figure 8.5 Losses for a 5-mil-wide (0.13-mm), 50Ω half-ounce copper stripline on FR4 as computed by Linpar [1]. Total loss in dB (αt) is the sum of the dielectric (αd) and conductor (αc) losses.
losses. This comes about because at a given frequency the dielectric loss in stripline
traces is the same for all widths, but it falls slightly as the width of microstrips
increase.
8.4 How Do Impedance and Loop Resistance Affect Conductor Loss?
The conductor loss (αc) in decibels (dB) is related to the AC loop resistance (Rac,
introduced in Chapter 7) and the trace impedance (Zo) as shown in (8.3), which is
valid for either stripline or microstrip traces:
αc =
4.3 × Rac
Zo
(8.3)
Because of the way in which field lines become concentrated at sharp corners,
calculating Rac by hand is not easy for rectangular traces. Fortunately, there are
84
Signal Loss and the Effects of Circuit Board Physical Factors
0.05
0.08
0.10
0.13
Width (mm)
0.15 0.18
0.20
0.23
0.25
0.13
0.051
50Ω
0.12
0.047
0.11
0.043
0.10
0.039
70Ω
0.09
0.035
0.08
0.031
0.07
0.028
0.06
0.024
2
3
4
5
6
7
8
9
dB/cm
dB/inch
60Ω
10
Width (mils)
Figure 8.6 Total signal loss from Linpar [1] at 350 MHz for half-ounce rectangular stripline (solid
curve) and microstrip (dotted curve) on FR4.
ways to accurately calculate microstrip and stripline loss without first calculating
Rac, and we will see how in Chapter 17. The purpose of introducing (8.3) here is to
show how the loop resistance and impedance interrelate to affect conductor loss.
Equation (8.3) shows how increasing the trace impedance lowers the conductor loss and how increasing the AC loop resistance increases the conductor loss.
This makes sense; long, narrow, thin, heavily undercut traces should have a higher
resistance (and so conductor loss) than short, wide, thick, perfectly rectangular
traces (which have a lower resistance).
Chapter 7 showed how the skin effect causes the loop resistance to increase as
the square root of frequency. From (8.3) this means the conductor loss will also.
Equation (8.3) has another implication. For a given resistance (and therefore
trace width), high-impedance traces will have a lower conductor loss than lowerimpedance traces. For example, a 65Ω, 5-mil-wide trace will have a lower conductor loss than a 50Ω, 5-mil-wide trace.
The Problems show how to use (8.3) and Rac to estimate the conductor loss.
8.4.1 How Does Surface Roughness Increase Conductor Loss?
Manufacturers roughen one or both sides of the copper foil sheet so it better adheres to the epoxy substrate material (see Chapter 5). An example of surface roughness is shown in Figure 8.7.
Chapter 7 showed how the skin effect causes high-frequency currents to flow
on the trace surface, increasing its resistance. Any irregularities or imperfections
present on the copper surface forces the highest frequency currents to flow through
the peaks and valleys of the rough surface, further raising the resistance to those
8.4 How Do Impedance and Loop Resistance Affect Conductor Loss?
85
Figure 8.7 Cross-section of a 1-ounce stripline. The roughness of the trace surface is clearly evident.
(Courtesy of TTM Technologies, Inc. Used with permission.)
currents. This added resistance, created by the need to roughen the copper for adhesion, becomes pronounced for frequencies where the skin depth is comparable to
the average roughness (Ra) of the copper foil [2]. This is explored in the Problems,
where we find this typically occurs for frequencies of a few gigahertz and more.
For those currents the conductor loss is larger than predicted by the square root of
f skin effect model.
The impact of surface roughness (Ra) is seen in Figure 8.8 for a 50Ω, 5-milwide solder mask covered microstrip to be small at frequencies of a few hundred
megahertz and below, but its effects steadily increase at higher frequencies.
8.4.2 Surface Roughness in Perspective
In general, it is important to include surface roughness effects in simulations when
high frequencies are present, especially if the circuit board has low dielectric losses.
Dielectric losses are lowest when the loss tangent is small and, in general, when using wide microstrips. As we have seen, stripline dielectric losses are usually higher
than microstrips for given impedance and trace width.
However, the impact of surface roughness on the total loss depends on the
relative magnitude of the dielectric loss. For instance, at 5 GHz (which represents
the fundamental frequency of a 10-Gbps binary NRZ signal stream), moderate
amounts of surface roughness (Ra = 1 μm) increase the microstrip conductor loss
by 8% more than a smooth trace, when but when dielectric loss for FR4 is included, the overall total loss is only increased by about 5%.
8.4.3 What Can Be Done to Reduce Conductor Losses?
To lower conductor losses:
•
Use wide traces since this lowers the loop resistance (Rac).
86
Signal Loss and the Effects of Circuit Board Physical Factors
1.0
0.39
0.35
0.8
Ra = 2 μm
Ra = 1 μm
0.7
Smooth
0.28
0.31
0.6
0.24
0.5
0.20
0.4
0.16
0.3
0.12
0.2
0.08
0.1
0.04
0.0
Attenuation (dB/cm)
Attenuation (dB/inch)
0.9
0.00
1 GHz
2 GHz
3 GHz
4 GHz
5 GHz
Frequency
Figure 8.8 Conductor loss of a 5-mil-wide, 50Ω, half-ounce solder mask covered microstrip with
and without surface roughness. A 60° tooth angle is applied to both trace and return plane on
adjacent surfaces. Resistance increases slightly for smaller angles. (Data from the Simbeor 3D electromagnetic field solver [3].)
•
When the traces are controlled impedance and the board is a fixed thickness,
select a laminate system with a lower dielectric constant. The Problems show
how this can sometimes improve conductor loss.
•
Specify high-impedance traces. This increases the distance to the return plane
(especially for wide traces), lowering the loop resistance.
•
Use copper with low surface roughness.
Often 50Ω traces are required for high-speed signaling, which can make it
impractical to lower conductor losses by signaling with higher impedance traces.
Note that because of the skin effect, much of the high-frequency current flows
near the trace surface, so increasing the copper thickness will not significantly reduce high-frequency conductor losses.
8.5
Understanding Dielectric Losses
In Chapter 7 we saw how conductance (the circuit element used to represent dielectric loss in many circuit simulators) is affected by the impedance and trace width
for microstrips and striplines. Here we examine dielectric losses directly, without
using conductance.
The amount of signal energy lost to the dielectric is affected by frequency, the
dielectric constant, and the loss tangent as shown in (8.4), which is slightly modified from that appearing in [4, 5].
8.5 Understanding Dielectric Losses
87
αd = K × f × εr × LT
(8.4)
In (8.4) αd is the dielectric loss at the frequency f (in gigahertz), and LT is the
loss tangent. To get the loss in dB per inch, set K to 2.32; use 0.91 to get the loss
in dB per cm. For stripline, εr is the dielectric constant of the laminate and prepreg,
but it can be replaced with the effective dielectric constant (εr_eff) for microstrip.
Equation (8.4) shows how the dielectric loss increases with frequency and the
loss tangent. The loss at 3.5 GHz will be 10 times the loss at 350 MHz (assuming
that the loss tangent remains the same). It also shows how small changes to the
loss tangent only slightly affect the dielectric loss. This is important because the
loss tangent for many high-performance laminate systems is only moderately lower
than typical FR4 values. This means that switching from FR4 to a higher-cost alternate laminate system will not reduce the total loss unless the signaling speeds
are very high.
8.5.1 Differences in Dielectric Losses Between Stripline and Microstrip
Because in stripline the dielectric constant is the same for all trace widths and impedances, (8.4) shows that losses in the dielectric are not affected by stripline geometry. For example, the dielectric loss for a 5-mil-wide, 50Ω stripline will be the same
as the losses for a 10-mil-wide 65Ω stripline fabricated on the same routing layer.
However, we have already seen how in microstrips the width and height above
the ground plane strongly affect the effective dielectric constant (εr_eff). It is evident
by using the effective dielectric constant in (8.4) that dielectric losses are affected
by the geometry of a microstrip. This means that because the geometry also affects
the impedance, microstrips of different impedances will have different dielectric
losses. In stripline the dielectric constant is the same of all impedance values, so the
dielectric losses are the same for all impedances.
This significant observation is illustrated in Figure 8.9.
Since air has a loss tangent lower than the laminate, dielectric losses will be
smaller for those microstrips that have proportionally more of the signal energy
propagating in air. This corresponds to wide, high-impedance microstrips because
they are further away from the ground plane than narrow, low-impedance ones.
It is for this reason that the dielectric losses are lower for wide, high-impedance
microstrips than narrow, low-impedance ones.
8.5.2 Effects of Temperature and Moisture
The dielectric constant and loss tangent values for laminate systems are affected by
temperature, humidity (moisture uptake), and the glass-to-resin ratio.
For instance, increasing the temperature from 25°C to 80°C causes the dielectric constant of typical FR4 epoxies to increase by under 10%, and the loss tangent
to fall by more than 40% [6]. From (8.4) this change is significant and the actual
operating temperature of the system should be factored into the loss model.
Exposing the FR4 laminate to 90% relative humidity for 200 hours can cause
the loss tangent to increase by up to 35% compared to its value when dry, and,
depending on the glass-to-resin ratio, the dielectric constant can increase by more
than 10% [5–7].
88
Signal Loss and the Effects of Circuit Board Physical Factors
Width (mm)
0.05
0.060
0.10
0.15
0.21
0.26
0.31
0.36
0.024
0.022
Stripline
0.050
0.020
50Ω Microstrip
65Ω Microstrip
0.045
Dielectric Loss (dB/cm)
Dielectric Loss (dB/inch)
0.055
0.018
0.040
0.016
2
4
6
8
10
12
14
Width (mils)
Figure 8.9 Dielectric loss of stripline is the same for all trace widths and impedances, but changes
for solder mask–covered microstrip. (Data for FR4 (εr = 4.2, LT = 0.02) from Linpar [1].)
8.5.3 What Can Be Done to Reduce Dielectric Losses?
To lower dielectric losses:
•
Use a low loss tangent laminate system.
•
Use a low dielectric constant laminate system.
•
Use wide, high-impedance traces when using microstrip since that lowers
εr_eff.
8.6 Summary of Signal Loss and Distortion Characteristics
•
A truly lossless line is distortionless. Provided the line is properly terminated,
pulses sent down these transmission lines do not change shape.
•
Signal energy is lost to the conductor resistance and to the dielectric.
•
A lossy line distorts the pulse by attenuating the high-frequency harmonics
making up the pulse.
•
Loss reduces the amplitude and rounds the pulse.
•
Conductor losses increase as the square root of frequency, but dielectric
losses increase directly with frequency.
•
The crossover frequency where the dielectric loss outweighs the conductor
losses depends on the trace width, its thickness, and its impedance.
8.7 Effects of Resin Content and Glass Weave on Delay Time and Impedance
•
89
For FR4 systems higher temperatures increase conductor resistance (and
therefore loss), while the temperature lowers loss tangent (and therefore dielectric losses).
8.7 Effects of Resin Content and Glass Weave on Delay Time and
Impedance
Researchers [8] have observed a variation in the dielectric constant of nearly 9%
occurring across a wide bus even when all its members are routed on the same layer.
This occurs because a trace running directly on top of a glass yarn bundle in
the laminate has a higher dielectric constant than one running along a resin filled
trough where there is no glass. Since the pitch of the glass yarn making up the mats
is typically between 16.7 and 23 mils (0.4–0.9 mm) [9], this effect is most likely to
be observed with wide (multitrace) buses, but it also can be significant for highspeed serial signaling involving only two traces [10].
For instance, a 16-bit bus routed with 5-mil-wide traces and 5-mil space between traces will occupy 155 mils (0.4 cm) in width, which is several times the
pitch of the glass yarn. It is very likely that at least one of the traces will route over
a glass bundle while another in the same bus routes over an epoxy trough.
This effect can be ignored when timing skew between various signals does not
have to be tightly controlled. When signaling at high speeds, the two extremes are
addressed by creating two transmission line models: one when the trace is routed
over glass and the other when routed over epoxy.
8.8 Effects of Trace Shape
Figures 5.5 and 5.7 show that actual circuit board traces are trapezoidal rather than
perfectly rectangular in shape.
Because the surface area exposed to the return path will be less, a trapezoidal
trace will have lower capacitance and higher inductance than rectangular traces,
which increases the trace impedance. Trapezoidal traces also have less copper to
support current flow, which increases their resistance and the amount of conductor
loss.
Using a perfectly rectangular 5-mil-wide, 50Ω stripline as an example, Rac increases by some 22%, but the inductance only increases by about 4% and the
capacitance falls by about 2.5% when the top of the trace is reduced to 3.7 mils
(corresponding to a heavily overetched trace). Because the impedance changes as
the square root, this only results in the impedance increasing to 51Ω.
8.9
Main Points
•
Loss grows with frequency, causing a pulse to be distorted in shape and reduced in amplitude as each of its harmonics is attenuated differently.
90
Signal Loss and the Effects of Circuit Board Physical Factors
•
Signal energy is lost to the trace resistance and to leakage currents in the
dielectric.
•
At low frequencies the conductor loss outweighs dielectric loss, but this reverses at high frequency.
•
High-impedance traces have lower conductor loss than low-impedance
traces.
•
Surface roughness of the trace and the return planes increase the conductor
loss, especially at high frequency.
•
Dielectric loss increases with frequency and is determined by the laminate’s
loss tangent value.
•
Changes in temperature alter the conductor and dielectric losses, and the
dielectric loss is also affected by moisture.
•
Traces are usually not rectangular in shape. This alters the transmission line’s
electrical characteristics.
Problems
Answers to these problems are available on the Artech House Web site at http://
www.artechhouse.com/static/reslib/thierauf/thierauf1.html.
8.1
8.2
8.3
8.4
8.5
8.6
8.7
A serial interface requires a minimum signal of 175 mV peak to peak at
the receiver. Assuming that the transmitter launches 800 mV peak to peak,
what is the maximum loss allowed in the interconnect?
Transmission line loss is a concern in a design linking two custom-designed
integrated circuits. What impedance and trace width should be selected to
minimize loss?
Estimate the conductor loss of a 15.75-inch (40-cm)-long, 8-mil (0.20-mm)wide, 50Ω half-ounce stripline on FR4 at 1.25 GHz.
In a particular application, 50Ω controlled impedance stripline traces on
FR4 have too much loss. The signal integrity engineer states that although it
will cost more to manufacture, both the dielectric and conductor losses can
be reduced by changing to a low Dk laminate. Is this true, and if so, why?
To reduce cost and current consumption, a gate array manufacture has
moved an old product line to a new process technology. A side effect is that
the I/O signal rise time that was 1 nS becomes 700 pS. According to the
manufacturer, this is advantageous because it results in more setup time for
downstream logic. Considering only signal integrity, what must be considered before you can decide if it is safe to use this new version of the chip in
place of the older chip in existing designs?
A 50Ω stripline has a dielectric loss of 3 dB at 2 GHz. What is the dielectric
loss at 1 GHz for a 100Ω stripline that is fabricated on the same circuit
board?
What the dielectric loss for a 40-cm-long stripline trace at 1 GHz when
fabricated on IS410?
Problems
8.8
8.9
91
A 4-ns pulse has harmonics every 250 MHz. Calculate the total loss for a
5-mil-wide, 50Ω stripline on FR4 for the pulse fundamental frequency and
the first five harmonics.
Determine the frequency where surface roughness is important to conductor loss.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
Djordjevic, A. R., et al., LINPAR for Windows, Norwood, MA: Artech House, 1999.
Brist, G., et al., “Non-Classical Conductor Losses Due to Copper Foil Roughness and
Treatment,” ECWC 10/IPC/APEX Conf., Fall 2005.
Simbeor Electromagnetic Simulation Environment, Simberian Corporation.
Cohn, S. B., “Problems in Strip Transmission Lines,” IEEE Transactions on Microwave
Theory and Techniques, Vol. MTT 3, No. 2, March 1955, pp. 119–126.
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Khan, S., “Comparison of the Dielectric Constant and Dissipation Factors of Non-Woven
Aramid/FR4 and Glass/FR4 Laminates,” Technical Note, DuPont Advanced Fibers Systems
Division, Richmond, VA, undated.
Matsushita Electronic Materials, Inc. “Megtron R5715 Technical Brochure,” Forest Grove,
OR, Number 199707-1.5Y, 1997.
Heck, H., et al., “Impact of FR4 Dielectric Non-Uniformity on the Performance of MultiGb/s Differential Signals,” IEEE Conf. Electrical Performance of Electronic Packaging,
October 27–29, 2003, pp. 243–246.
Brist, G., et al., “Woven Glass Reinforcement Patterns,” Printed Circuit Design & Manufacture, November 2004, pp. 28–33.
McMorrow, S., and C. Heard, “The Impact of PCB Laminate Weave on the Electrical Performance of Differential Signaling at Multi-Gigabit Data Rates,” DesignCon 2005, Santa
Clara, CA, 2005.
CHAPTER 9
Understanding Trace-to-Trace Coupling
9.1
Introduction
This chapter discusses how coupling caused by parasitic capacitance and inductance linking traces changes their impedance and propagation time. We will use this
understanding in the following chapters, especially when discussing crosstalk, serial
signaling, and transmission line terminations and reflections.
9.2 Understanding Mutual Capacitance and Inductance
Chapter 7 showed how capacitance is created between a trace and the ground plane
(the return path). In fact, as shown in Figure 9.1(a), besides this self-capacitance
(and implied but not shown, self-inductance), parasitic coupling capacitance and
inductance are also present between traces.
We know from Chapter 7 that a transmission line can be thought of as a chain
of very small, identical distributed circuit elements. It is apparent from Figure
9.1(b) that this is also true for the parasitic circuit elements between traces (the mutual terms). The line itself is made from an infinite number of these tiny segments,
each with their own identical self and mutual elements. This means that similar to
the self terms, the mutuals can be described per unit length (PUL), introduced in
Chapter 6).
9.2.1 What Are the Capacitance and Inductance Matrices?
It is convenient (and mathematically desirable) to record all of the self and mutual
elements as matrices.
As shown in (9.1a) and (9.1b), and in Figure 9.1, a two-digit numbering scheme
is used to identify the terms.
L11
L = L21
L31
L12
L22
L32
L13
L23
L33
(9.1a)
93
94
Understanding Trace-to-Trace Coupling
C 13
1
2
C 12
C 23
C 30
C 20
C 10
3
Ground
plane
L 11
Trace 1
C 10
L 13
L 12
C 12
L 22
Trace 2
C 20
C 13
L 23
C 23
L 33
Trace 3
C 30
Figure 9.1 (a) Trace cross-section and (b) circuit model of a small segment of line showing the
capacitances and inductances for three traces.
C11 C12
C = C21 C22
C31 C32
C13
C23
C33
(9.1b)
The self-inductance (Ls) of trace 1 is L11 and is measured with no current flowing in the other traces. The mutual inductance (Lm) linking trace 1 to 2 is L12.
The self-capacitance (Cs) and mutual capacitance (Cm) are similar, but as shown
in (9.2), C11 and the other diagonal terms such as C22 in the matrix is the capacitance to ground (C10 in Figure 9.1) plus the sum of all the mutual capacitances
when they are grounded.
C11 = C10 + C12 + C13
(9.2)
By rearranging (9.2), we find that the capacitance to ground of just trace 1
(C10) is the self-capacitance (C11) minus all of the mutual capacitance.
To illustrate, the inductance and capacitance matrices per meter for a pair of
traces are found by a 2D field solver to be:
2.876E-07 3.564E-08 9.859E-09
L = 3.564E-08 2.863E-07 3.564E-08
9.859E-09 3.564E-08 2.876E-07
(9.3a)
9.3 How Does Switching Alter the Trace Inductance and Capacitance?
1.176E-10 −5.306E-12 −6.049E-13
C = −5.306E-12 1.180E-10 −5.306E-12
−6.049E-13 −5.36E-12 1.176E-10
95
(9.3b)
In this example, the self-inductance (L11) is 287.6 nH/m, the mutual inductance from trace 1 to 2 (L12) is 35.64 nH/m, and the mutual inductance from trace
1 to 3 (L13) is 9.859 nH/m.
The total capacitance to ground on trace 1 (C11) is 117.6 pF/m when traces
2 and 3 are grounded. The mutual capacitance from trace 1 to 2 (C12) is −5.306
pF/m; it is −0.6049 pF/m from trace 1 to 3 (C13). In the following sections we will
see the reasoning and benefits of showing the mutual capacitances as negative numbers. Here we notice that the total mutual capacitance is the sum of the two mutual
terms: −5.911 pF.
This makes intuitive sense. Because traces 1 and 2 are closer together than
traces 1 and 3, the mutual capacitance between traces 1 and 2 should be higher
than between traces 1 and 3, and since the self-capacitance includes all the mutual
capacitance, C11 must be higher than the value of the mutual terms. This is further
explored in the Problems.
By examining these matrices, we notice that the mutual terms on either side
of the diagonal are the same. For instance, L12 has the same value as L21 (35.64
nH/m). This is reasonable because we would expect the coupling between traces
1 and 2 to be the same whether it is measured from traces 1 to 2, or from traces 2
to 1. SPICE-type simulators take advantage of this symmetry and require that only
half of the matrix be entered when creating transmission line models.
9.2.2 Why Do Some Field Solvers Report Capacitances with Negative
Numbers?
Field solvers commonly report the mutual capacitances as negative numbers, but
this is not a universal practice. The negative value for the mutual capacitance is
mathematically correct and is an artifact of the way in which charge is calculated
when the field solver computes the capacitance of a structure. Those interested in
these details are referred to [1–3].
Although the negative values are proper, their presence is a common source of
confusion because all of the hand calculations treat the mutual capacitance values
as positive numbers.
As shown in the Problems, calculations throughout this book always take the
mutual capacitance as positive, even if the field solver reports them as negative.
9.3 How Does Switching Alter the Trace Inductance and Capacitance?
Figure 9.2 shows a trace flanked by neighboring traces. The aggressors switch at the
same time as the victim trace, and all of them have the same signal amplitude. This
implies that the voltage on each trace has the same rate of change. In practice loading differences may make the signal rise times unequal, but to simplify the analysis
we will assume that they are the same.
96
Understanding Trace-to-Trace Coupling
Aggressor Trace 1
Victim Trace
Aggressor Trace 2
Figure 9.2 Top view showing a victim trace flanked by neighboring aggressor traces.
9.3.1 How Is the Capacitance Affected?
The capacitance for a tiny segment of the three traces illustrated in Figure 9.1 and
9.2 appears in Figure 9.3.
Figure 9.3 shows that when the victim and aggressors simultaneously switch in
phase with identical rise times, the voltage at each terminal simultaneously reaches
1V. This makes the voltage difference across either of the mutual capacitors (C12 or
C23) 0, which means no current flows and these capacitors can be removed without
altering the circuit behavior. The net result is that the total switched capacitance of
trace 2 is only C20: its capacitance to ground.
0 → 1V
Aggressor 1
C 12
ΔV = 0V
1V
0 → 1V
Victim
C 20
C 23
0 → 1V
Victim
C 20
ΔV = 0V
Aggressor 2
Figure 9.3 When the victim and aggressor traces switch in phase, the victim capacitance is the
self-capacitance to ground (C20).
9.3 How Does Switching Alter the Trace Inductance and Capacitance?
97
The situation is quite different when then signals switch out of phase. Figure
9.4 shows the victim switching low to high while the aggressors surrounding it
switch in the opposite direction.
In this case the aggressors switch from 1V to 0V while the victim switches from
0V to 1V. The mutual capacitances C12 and C23 experience a 2V change in voltage,
effectively multiplying their value by a factor of 2. As shown in Figure 9.4, this
makes the total capacitance on the victim trace equal to twice the mutual capacitance plus C20 (the capacitance of the victim trace to ground).
From (9.2) the victim’s self-capacitance (in this case, C22) already includes the
mutual capacitance. As shown in Figure 9.4, this allows us to simplify the total capacitance to be the sum of the self-capacitance (C22) plus the mutual capacitances.
9.3.2 How Is the Inductance Affected?
From circuit theory (for example, see [4]), the mutual inductance occurring between
side-by-side wires can either increase or decrease the total circuit inductance depending on the direction of the current flow. In particular, when the signals switch
out of phase, the currents in the aggressor lines are flowing in the opposite direction to the current in the victim, and the mutual inductance subtracts from the
self-inductance.
However, the currents flow in the same direction when the signals switch in
phase. In this situation the mutual inductances add to the self inductances. This
useful observation will be used later in this chapter when we analyze switching
modes.
1 → 0V
Aggressor 1
C12
ΔV = 2V
1V
0 → 1V
Victim
C 20
C 23
1 → 0V
Victim
ΔV = 2V
C 22 + C 12 + C 23
Ctotal = C 20 + 2C 12 + 2C 23 = C 22 + C 12 + C 23
Aggressor 2
Figure 9.4 When the victim and aggressor traces switch out of phase, the victim capacitance is the
capacitance of C2 plus the mutual capacitances (C22 + C12 + C23).
98
Understanding Trace-to-Trace Coupling
9.4 What Effect Does Coupling Have on Impedance and Delay?
The circuit setup shown in Figure 9.5 shows four aggressor microstrip traces (A4
through A1) surrounding a victim (V) microstrip. They are designed to have an
impedance of 50Ω, and are driven from ASIC 1 with identical I/O drivers.
By systematically switching each aggressor in or out of phase with the victim
trace, we can determine how the mutual capacitance and inductance affect the
victim impedance and delay for each switching mode. This is shown in Table 9.1,
where a “+” indicates that the signal transition is in the same direction, while a “−”
means the signals swing in opposite directions. A signal that does not switch (is
static) is shown with an S.
Evidently, even though the traces are designed to be 50Ω, they only have that
value when their neighbors are static. The impedance is lowest when all of the
aggressors are out of phase with the victim, and it is highest when they are all in
phase.
The switching activity also affects the microstrip (but not stripline) propagation time. Microstrip delay is shortest when the signals are all out of phase, and it
is the longest when they are all in phase. As we saw in Chapter 7, stripline propagation delay is determined by the dielectric constant, which is why it is the same for
all switching modes and so is not shown in Table 9.1.
However, data-dependent timing jitter will occur at the output of receivers connected to either stripline or microstrip traces. This behavior is a property of the circuit rather than the transmission line and is an artifact of the changing impedance
ASIC 2
ASIC 1
1m
A4
tpd
A3
5 mils
V
A2
5 mils
A1
Figure 9.5 Top view of four aggressor microstrip traces surrounding a victim trace connecting ASIC
1 to ASIC 2 and driven by identical drivers.
9.5 What Are Odd and Even Modes?
99
Table 9.1 Impedance and propagation delay
of the victim 50Ω microstrip in Figure 9.5.
A4
A3
V
A2
A1
Z
tpd
−
−
+
−
−
42
6.31
+
−
+
−
+
43
6.43
S
S
+
S
S
50
6.46
−
−
+
+
+
52
6.54
−
+
+
+
−
63
6.65
+
+
+
+
+
69
6.92
Transmission line impedance (Z) is in ohms and the delay (tpd)
is in nanoseconds/meter.
of the victim trace causing different voltages to be launched when neighboring
traces switch.
To summarize:
•
Stripline and microstrip victim traces have the nominal impedance when all
neighbors are static.
•
The lowest impedance for stripline or microstrip victims occurs when all
neighbors switch out of phase.
•
The highest impedance of stripline or microstrip victims occurs when all
neighbors switch in phase.
•
Stripline propagation delay is not affected by the switching activity of its
neighbors.
•
Microstrip has the nominal delay when all its neighbors are static.
•
Microstrip has the lowest propagation time when all its neighbors switch out
of phase and has the largest propagation time when all its neighbors switch
in phase.
9.5 What Are Odd and Even Modes?
In a two-trace system the switching mode is said to be even when the traces switch
in phase; it is odd when they are out of phase [5, 6].
Differential signaling (as discussed in Chapter 13, where one trace carries a
positive going signal and the other carries its complement) is an example of oddmode switching since the two signals are always intended to be out of phase.
From Figure 9.6 with even-mode switching, the inductive coupling moves in
the same direction and encompasses both traces, while the capacitive coupling does
not include the neighboring traces. From this we would expect the total inductance
to be equal to the self plus the mutual inductance and the total capacitance to be
only the capacitance to ground. Since by its definition the self-capacitance includes
all of the mutual capacitance, to obtain the total even-mode capacitance, the mutual capacitance must be subtracted from the self-capacitance.
100
Understanding Trace-to-Trace Coupling
+
+
+
−
+
Even mode
switching
(a)
Odd mode
switching
(b)
Figure 9.6 Edge view of two pairs of traces switching in (a) even mode and (b) odd mode. Electric
fields (capacitive coupling) are shown as solid lines and magnetic fields (inductive coupling) are
shown as dotted lines.
This contrasts with odd mode where the magnetic field lines move in opposite
directions and the electric fields extend from one trace to the other. From this we
would expect the inductance to only include the self-inductance and the capacitance to include both the self- and mutual capacitances.
This reasoning is confirmed in (9.4) and (9.5) for even-mode switching and in
(9.6) and (9.7) for odd-mode switching [5]. These equations may be used for either
microstrip or striplines.
Zoe =
tpdeven =
(Ls + Lm )(Cs − Cm )
Zoo =
tpdodd =
Ls + Lm
C s − Cm
Ls − Lm
C s + Cm
(Ls − Lm )(Cs + Cm )
(9.4)
(9.5)
(9.6)
(9.7)
The even-mode impedance is called Zoe and odd-mode impedance is called Zoo
to distinguish them from the characteristic impedance (Zo). Similarly, tpd_even and
tpd_odd distinguish the even- and odd-mode delays from the propagation time of a
single trace.
In the equations Ls and Cs are the self-inductance and capacitance as displayed
by a field solver, and Lm and Cm are the sum of all the mutual terms. All are per
unit length. For instance, from the capacitance matrix in (9.3b), Cs is 117.6 pF/m
and Cm is −5.91 pF/m.
We will return to these equations in future chapters. Here we again note that
the mutual capacitance is taken as positive in the equations, even if they are reported as negative. As shown in the Problems, the even-mode impedance for trace
1 using the information in (9.3) is 54.6Ω, and not 51.9Ω because +5.9 pF is used
for Cm, even though it is reported as −5.9 pF.
9.5 What Are Odd and Even Modes?
101
9.5.1 Can the Odd- and Even-Mode Equations Be Used with More Than Two
Traces?
As demonstrated in the Problems, by summing all of the mutuals and using them
as Lm and Cm, (9.4) through (9.7) can be used to find the impedance and delay of a
victim trace when many neighbors switch.
9.5.2 Why Is the Even- and Odd-Mode Timing the Same for Stripline But Not
for Microstrip?
From Chapter 7, we know that in stripline the propagation delay is determined by
the laminate dielectric constant and not by the trace shape or impedance. This appears to conflict with (9.5) and (9.7) because they show that, in general, the timing
for even and odd modes is different.
We must turn to field theory to resolve this apparent paradox. It can be shown
[7, 8] that for the types of transmission lines of interest to us (lines where the electric and magnetic fields travel crosswise to the direction in which the signal is moving, called transverse electromagnetic, or TEM, transmission), the even-mode and
odd-mode propagation delays are the same. The significance of this is that because
striplines with low loss are true TEM lines, the even-mode propagation delay given
by (9.5) will be the same as the odd-mode propagation delay given in (9.7). In fact,
when used with striplines, they will give the same result as (7.5) where we used the
dielectric constant to find the delay. This is summarized in Figure 9.7 and explored
in the Problems.
However, the situation is very different in microstrips because even when they
are lossless, the two or more dielectrics involved means the signal propagation is
not truly TEM. In fact, microstrip delay is determined by how much of the signal
energy propagates in air and how much propagates in the laminate. This makes the
delay shorter when the victim and aggressors are out of phase because under those
conditions more of the signal energy travels in the air than in the laminate. The
Problems illustrate this further by showing that the odd-mode microstrip propagation time is smaller than the even-mode propagation times.
Stripline
Embedded
microstrip
tpd_odd = tpd_even
tpd_odd < tpd_even
Zoo < Z oe
Zoo < Z oe
Figure 9.7 Odd- and even-mode delays are the same in stripline traces but not in solder mask–covered microstrips.
102
Understanding Trace-to-Trace Coupling
9.5.3 By How Much Does the Even- and Odd-Mode Impedance Change?
We saw in Table 9.1 how coupling between traces during switching changes the victim trace impedance. The coupling can be reduced by increasing the spacing, which
makes the even- and odd-mode impedances equal. This is confirmed in Figure 9.8,
which shows the odd- and even-mode impedances with respect to the edge-to-edge
spacing for a pair of 5-mil-wide, half-ounce 50Ω and 65Ω stripline and microstrip
without solder mask traces on FR4.
The difference between the even- and odd-mode impedances is greatest when
the separation is small and the impedance is high, since those conditions make
coupling the highest. Increasing the trace-to-trace spacing reduces the coupling and
causes the even-mode impedance to fall and the odd-mode impedance to increase.
Eventually the traces are far enough apart to where they both equal the value the
traces would have in isolation.
Graphs such as Figure 9.8 are used to judge the relative coupling between a
pair of traces. For instance, we see the 50Ω traces have a reasonably loose coupling
(indicated by the small difference between the even- and odd-mode impedances)
with an edge-to-edge separation of 20 mils. However, the impedance difference is
quite a bit larger for the 65 traces. This shows that even though the traces are the
same width, for a 20-mil spacing the 50Ω traces are loosely coupled, but the 65Ω
traces are not.
This observation highlights the danger of applying simple rules of thumb to
determine if coupling is severe between traces. For instance, the 3-W rule is sometimes used to determine how close together traces can be routed. This well-known
rule [10] is illustrated in Figure 9.9 and holds that to minimize coupling the center
distance between a pair of traces should be three or more times the trace width.
90
Impedance (Ω)
80
Stripline
70
Microstrip
Even mode
Odd mode
60
50
40
0
10
20
30
40
50
Edge-to-edge spacing “s” (mils)
Figure 9.8 Odd- and even-mode impedance for a 5-mil-wide, half-ounce 50Ω and 65Ω stripline
(solid curve) and microstrip (dotted curve) on FR4. (From: [9]. Used with permission.)
9.6 What Are the Circuit Effects When Switching Causes the Impedance to Change?
103
3W
W
Figure 9.9 Side view of two traces showing the 3W rule of thumb.
Using this rule with 5-mil-wide traces would result in traces being routed as
close together as 10 mils. Judging from Figure 9.8, for the 65Ω trace the coupling
is quite high at that distance and it must be increased to over 25 mils before the
coupling is reduced to a reasonable value.
9.6 What Are the Circuit Effects When Switching Causes the
Impedance to Change?
Chapter 6 showed how voltage divider action between the impedances of the trace
and driver determines the voltage launched down the victim trace. Because switching activity of neighboring traces causes the victim impedance to change, neighbors
switching causes the launched voltage to change even if the power supply voltage
is held constant.
This is illustrated in Figure 9.10, which shows the voltage launched down the
transmission line circuit appearing in Figure 9.5, which was used to create Table
9.1. The drivers have a drive strength of 32 mA.
Figure 9.10 illustrates the two worst-case situations that bound the results.
Even though the power supply is 3.3V, the voltage launched by a 32-mA driver is
3.1V when the victim and neighbors are all switching in phase (corresponding to
a victim impedance of 69Ω in Table 9.1), but it falls to 2.9V when the neighbors
switch out of phase (the 42Ω impedance case). As explored in the Problems, a
weaker driver would launch a lower voltage and the difference in launched voltages
would be greater then the 200 mV shown in Figure 9.10.
9.7 How Is Receiver Timing Affected by In- and Out-of-Phase
Switching?
Table 9.1 shows how the switching activity of neighboring microstrips causes the
trace delay to vary from a low of 6.31 ns to a high of 6.92 ns, for a total variation
of 610 ps. However, as Figure 9.11 illustrates, the variation is 650 ps when the
measurement is made at the output of the victim receiver.
The added delay is caused by the difference in launched rise times between the
in-phase and out-of-phase cases. With an impedance of 69Ω, the in-phase signaling case represents less capacitive loading to the driver and so has a sharper rise
time than the out-of-phase case, which has an impedance of 42Ω. The result is that
more time is required for the signal to reach the receiver switch point. This loading
104
Understanding Trace-to-Trace Coupling
3.1V
3V
2.9V
2V
1V
0V
0.0s
5.0 ns
10.0 ns
15.0 ns
20.0 ns
Figure 9.10 Launched voltage is lower when the signals are out of phase (dotted curve) than when
in phase (solid curve).
3V
2V
650 ps
1V
0V
0.0s
5.0 ns
10.0 ns
15.0 ns
20.0 ns
Figure 9.11 Victim receiver responds sooner when the traces switch out of phases (dotted curve)
than when they are in phase (solid curve).
behavior is characteristic of CMOS I/O drivers and is independent of any parasitic
inductance in the power or signal leads.
9.8
9.8
Main Points
105
Main Points
•
Inductive and capacitive coupling is present between traces.
•
The switching activity of neighboring aggressor traces determines the victim
trace impedance and for microstrip (but not stripline) the propagation delay.
•
When only two traces are involved, the switching patters are called even
mode and odd mode.
•
Odd mode occurs when the victim and aggressor switch out of phase. This
gives the lowest impedance and for microstrip the shortest propagation delay.
•
Even mode occurs when the victim and aggressor switch in phase. This gives
the highest impedance and for microstrip the longest propagation delay.
•
The even- and odd-mode delay is the same for stripline.
•
The change in impedance caused by switching modes affects the launched
voltage for both microstrip and stripline traces.
Problems
Answers to these problems are available on the Artech House Web site at http://
www.artechhouse.com/static/reslib/thierauf/thierauf1.html.
9.1
9.2
9.3
Under switching conditions a victim trace in a two-trace system has the following impedance values: 36Ω, 60Ω, and 97Ω. Identify the odd, even, and
characteristic impedances.
Under switching conditions a victim trace in a two-trace system has the following propagation times: 154 pS/inch and 142 ps/inch. Identify the odd
and even propagation delays.
A field solver produced the following output file for a five-trace system.
Identify Ls, Lm, Cs, and Cm for trace 3.
L MATRIX [H/M]
2.875E-07
3.557E-08
9.705E-09
4.228E-09
2.416E-09
3.557E-08
2.862E-07
3.530E-08
9.620E-09
4.228E-09
9.705E-09
3.530E-08
2.861E-07
3.530E-08
9.705E-09
4.228E-09
9.620E-09
3.530E-08
2.862E-07
3.557E-08
2.416E-09
4.228E-09
9.705E-09
3.557E-08
2.875E-07
C MATRIX [F/M]
1.176E-10
−5.300E-12 −5.833E-13 −2.531E-13 −1.521E-13
−5.300E-12 1.180E-10
−5.261E-12 −5.677E-13 −2.531E-13
−5.833E-13 −5.261E-12 1.180E-10
−5.261E-12 −5.833E-13
−2.531E-13 −5.677E-13 −5.261E-12 1.180E-10
−5.300E-12
−1.521E-13 −2.531E-13 −5.833E-13 −5.300E-12 1.176E-10
9.4
Use the inductance and capacitance matrix from Problem 9.3 to find the
two worst-case impedance values for trace 3.
106
Understanding Trace-to-Trace Coupling
9.5
9.6
9.7
Use the inductance and capacitance matrix from Problem 9.3 to find the
two worst-case propagation times.
Does the matrix in Problem 9.3 represent a microstrip or stripline?
A field solver produced the following output file for a two-trace system. Is
this the matrix for a stripline or microstrip?
L MATRIX [H/INCH]
1.052e-8
2.040e-9
2.040e-9
1.052e-8
C MATRIX [F/INCH]
2.835e-12
−5.490e-13
−5.490e-13
2.835e-12
9.8
Compare the results from (9.5) and (9.7) with (7.5) for the two-trace system
given in Problem 9.7. Assume that the dielectric constant is 4.0.
9.9 Use the matrices in Problem 9.3 to show how the coupling falls as the distance from a given trace increases. Also show that the traces are symmetrical.
9.10 What voltage will be launched down the transmission line described by the
matrix appearing in Problem 9.7, assuming that the ASIC I/O driver has an
output impedance of 25Ω and is connected to a 2.5-V supply? How does
the voltage change when the driver impedance is reduced to 10Ω?
9.11 Use the matrices in Problem 9.7 to show that incorrect results are obtained
when negative values are used for mutual capacitance when calculating the
even- and odd-mode delays and impedances.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
Young, B., Digital Signal Integrity, Upper Saddle River, NJ: Prentice-Hall, 2001.
Paul, C. R., Analysis of Multiconductor Transmission Lines, New York: John Wiley &
Sons, 1994.
Baloglu, H. B., Circuits, Interconnections, and Packaging for VLSI, Reading, MA: Addison-Wesley, 1990.
Johnson, D., et al., Basic Electric Circuit Analysis, 5th ed., Upper Saddle River, NJ: Prentice-Hall, 1995.
Marx, K. D., “Propagation Modes, Equivalent Circuits, and Characteristic Terminations
for Multiconductor Transmission Lines with Inhomogeneous Dielectrics,” IEEE Transactions on Microwave Theory and Techniques, Vol. MTT-21, No. 7, July 1973, pp. 450–457.
Trinogga, G., et al., Practical Microstrip Circuit Design, Upper Saddle River, NJ: PrenticeHall, 1991.
Matick, R., Transmission Line for Digital and Communication Networks, New York: McGraw-Hill, 1969.
Ramo, S., et al., Fields and Waves in Communication Electronics, 3rd ed., New York: John
Wiley & Sons, 1994.
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Montrose, M. I., Printed Circuit Board Design Techniques for EMC Compliance, New
York: IEEE Press, 1996.
CHAP TE R 10
Understanding Crosstalk
10.1
Introduction
Crosstalk is the coupling of signal energy from one or more aggressor signals to one
or more victim signals. This causes unintended currents to flow in the victim, which
creates noise voltages.
When designers think of crosstalk, they often only consider how the noise voltage causes false triggering at a receiver and so disturbs system timing or reduces
noise margin. Signal integrity engineers take a wider view and recognize that in
addition to those types of problems, crosstalk is capable of causing pulse distortion
and (as described in Chapter 2) damage to the I/O cell if the transient voltages are
large enough.
10.2 How Is Crosstalk Created and What Are Its Characteristics?
The mutual capacitance and inductance appearing between the aggressor and victim traces [1–6] create the coupling responsible for crosstalk. This coupling was
introduced in Chapter 9 and is shown for two traces in Figure 10.1.
Although only two traces are shown, the mutual capacitance (Cm) and mutual
inductance (Lm) represent the sum of the electric and magnetic coupling from any
number of aggressors to the victim.
The electric and magnetic coupling causes current to flow in the victim [8], as
shown in Figure 10.2.
By making the loads resistors equal in value to the impedance of the line, Ohm’s
law can be used to find the load voltages, and reflections (discussed in Chapter 11)
are eliminated.
The current injected by the mutual capacitance (Ic) flows in the victim trace
with half flowing toward the near end and the other half flowing toward the far
end. The current induced by the mutual inductance (Il) does not behave in this way
because Lenz’s law causes the victim current to flow in the direction opposite to the
aggressor current (Is). For these things to be so, the victim and aggressors must be
far enough apart that the presence of the victim does not change the aggressors’ impedance (and so the voltages coupled to the victim cannot influence the aggressor).
107
108
Understanding Crosstalk
RL
RL
Separation
”s”
Ls
Lm
Aggressor
trace
FEXT
voltage
Ls
Victim
trace
Cm
Cs
Cs
RL
RL
NEXT
voltage
Figure 10.1 Electric and magnetic coupling between traces transfer energy from the aggressor to
victim trace. The mutual capacitance (Cm) represents the electric coupling, and mutual inductance
(Lm) represents the magnetic coupling. (From: [7]. Used with permission.)
IC
IC
Victim
FEXT = (I c − I L )RL
NEXT = (Ic + IL )RL
IL
RL = Zo
IL
Cm
Lm
Is
RL = Zo
Vs
RL = Zo
Is
Culprit
V L = R L Is
Is
RL = Zo
dv
dt
Length = l
Delay = tpd
Figure 10.2 Mutual capacitance and inductance cause current to flow in the tiny portion of the victim
trace shown. (From: [7]. Used with permission.)
The lines are considered to be loosely coupled in this way when the coupling is
under 25% [3].
The result is that the current from the mutual inductance flows back toward
the source, while the current from the mutual capacitance flows in both directions.
The currents combine to create a total noise current at each end of the victim.
At the near end the noise current is the sum of the two and creates the near-end
crosstalk voltage (NEXT). At the far end the two noise currents flow in opposite
directions, making the total the difference between the current coupled by the mutual capacitance and the current coupled by the mutual inductance. The combined
10.3
Far-End Crosstalk
109
current creates far-end crosstalk (FEXT) when it flows through the far-end load
resistance.
It is apparent from this discussion that FEXT and NEXT have very different
electrical properties, and in fact they must be analyzed separately.
10.3
Far-End Crosstalk
If the current from magnetic coupling is greater than the capacitive coupling current, the victim FEXT pulse will have a polarity opposite of the aggressor pulse.
The polarity will be the same if the magnetic coupling is less than the capacitive
coupling. If the magnetic and capacitive couplings are equal, the currents cancel
and there is no FEXT pulse. This is the case with stripline, provided that the losses
are low. However, because the currents never fully cancel, FEXT is always possible
in microstrips.
On circuit boards the coupling is usually mostly capacitive to nearby traces,
but it becomes increasingly inductive to traces progressively further away [1]. This
is why on circuit boards FEXT polarity is usually opposite of the aggressor pulse
polarity, but it can have the same polarity if the coupled distances are great. This
means that both the amplitude and the polarity of the FEXT pulse depend on the
switching behavior of neighboring traces.
As illustrated in Figure 10.3, the FEXT pulse has a width equal to the rise
time of the aggressor signal (dt, the change in time). This is because there is no DC
connection between the traces, so coupling only occurs while the aggressor pulse
is transitioning. This AC connection means that the FEXT pulse is no longer created once the aggressor pulse has reached its steady state value. In fact, FEXT is
determined by the change in voltage (dv) rather than the magnitude of the voltage,
and because the aggressor pulse has two edges, two FEXT pulses are created from
a single aggressor pulse.
10.3.1 Calculating FEXT
The forward crosstalk coupling factor (Kf) presented in (10.1a) shows the amount
and polarity of the coupling between traces. It has units of time per unit length (for
example, picoseconds/inch) and is used with (10.1b) and (10.1c) to find the actual
FEXT voltage. [4, 5, 8–10].
⎛L
⎞
Kf = −0.5 × ⎜ m − (Zo × Cm )⎟
⎝ Zo
⎠
FEXT = Kf × length ×
dV
dt
(10.1a)
(10.1b)
110
Understanding Crosstalk
FEXT = Kf × l × (dv/dt)
(Capacitive coupling)
Victim
tpd
dt
Victim
FEXT = Kf × l × (dv/dt)
(Inductive coupling)
dv
Aggressor
dt
Voltage
Time
Figure 10.3 FEXT victim pulse width equals the aggressor rise time (dt). It has the same polarity as the aggressor when capacitive coupling dominates but the opposite when inductive coupling dominates.
Zo =
Ls
Cs
(10.1c)
The sum of all the mutuals can be used in these equations to find Kf when more
than one trace aggresses a victim. The Problems include a worked example of how
to do this.
Because the mutual capacitance (Cm) term subtracts from the mutual inductance term (Lm), it is apparent from (10.1a) how the inductive and capacitive effects can cancel one another as happens with stripline. This is explored in the
Problems. It is also apparent how either the mutual inductance or capacitance can
become dominant to determine the polarity of Kf.
The change in voltage (dv) and time (dt) emphasize that it is the difference between voltage levels (rather than the magnitude) that creates FEXT. For instance, a
signal with a 1-ns rise time switching between 0.6V and 1.6V will induce the same
Far-End Crosstalk
111
FEXT amplitude as a 1-ns rise time signal switching from 0V to 1V since they both
have a 1V-per-nanosecond swing.
It is apparent from (10.1b) that the FEXT amplitude grows as Kf is made
larger, as the length of the coupling region increases, as the voltage swing (dv) increases, and when rise time (dt) is small. Therefore, FEXT is combated by lowering
the coupling between traces since that lowers Kf, keeping the coupled length short,
reducing the voltage swing of the aggressor, and lengthening its rise time.
10.3.2 What Are Typical Kf Values?
To find the forward crosstalk for a set of traces, a 2D field solver is used to determine the mutual capacitance and inductance, and Kf and then FEXT are found with
(10.1a) through (10.1c).
Figure 10.4 shows typical Kf values for a pair of 5-mil (0.13-mm)-wide, halfounce 65Ω and 50Ω solder mask–covered microstrips on FR4 (εr = 4.2). Since forward crosstalk is not present on low loss stripline, the graph only shows microstrip.
The separation is given as multiples of the trace width. A separation of 3 represents 15 mils since the trace is 5 mils wide (or 0.39 mm for a 0.13-mm trace).
We find from Figure 10.4 that Kf is −160 pS/m for a 5-mil (0.13-mm)-wide,
50Ω microstrip separated by 5 mils (“separation” of 1) from its aggressor. Using
this in (10.1b) and assuming that the launched pulse has a 2V swing (dv in the
equation) and a 1-ns rise time (dt), the forward crosstalk pulse is found to be −320
mV if the traces are coupled for a length of 1 meter. The crosstalk would be −160
mV if the coupled region was a half-meter in length.
0s/m
−20 ps/m
−40 ps/m
−60 ps/m
−80 ps/m
Kf
10.3
−100 ps/m
−120 ps/m
50Ω
65Ω
−140 ps/m
−160 ps/m
−180 ps/m
2
4
6
8
10
Separation (trace width multiples)
Figure 10.4 Kf for half-ounce, 5-mil (0.13-mm)-wide 50Ω and 65Ω solder mask–covered microstrips on FR4. (Raw data from the Polar Si9000 2D field solver [11].)
112
Understanding Crosstalk
Because the 65Ω trace is further away from the ground plane than the 50Ω
trace, more of the field lines can couple to the victim and less couple to the ground
plane. This makes Kf larger (more negative) for the 65Ω trace.
Because for these two cases inductive coupling outweighs capacitive coupling,
Kf is negative. It can become positive when the microstrip is covered with a thick
dielectric such as a heavy application of solder mask or conformal coating (provided that it has a higher dielectric constant).
10.3.3 FEXT Summary
To summarize FEXT behavior:
10.4
•
FEXT is 0 in striplines, provided that the transmission line is not very lossy.
FEXT is always possible in microstrips.
•
The FEXT pulse will have the opposite polarity of the aggressor pulse when
the magnetic coupling is greater than the capacitive coupling. It will have the
same polarity when the magnetic coupling is less than the capacitive coupling.
•
The FEXT pulse will have a higher amplitude if the aggressor rise time is
small or if the aggressor has a large change in voltage.
•
The FEXT pulse amplitude increases as the coupled length increases.
•
The FEXT pulse width is equal to the aggressor rise time.
Near-End Crosstalk
Referring to Figure 10.2, it is evident that in NEXT the inductively and capacitively
coupled currents are flowing in the same direction and add when they reach the
load at the near end. For this reason the NEXT pulse induced on the victim line
always has the same polarity as the aggressor pulse.
Because the currents flow away from the aggressor pulse, the current is continually induced in the victim as long as the aggressor pulse travels down the line.
At each instant as it moves, the aggressor pulse reaches a fresh mutual capacitor
and inductor, allowing current to be continually injected into the victim trace.
As illustrated in Figure 10.5, if the coupled region is electrically long (that is,
if the aggressor rise time is shorter than twice the delay of the coupled region),
the net result is a NEXT pulse having a width equal to twice the electrical length
of the coupled line, and an amplitude that saturates at a value determined by the
aggressors change in voltage. However, if the coupling region is short, the NEXT
amplitude cannot reach its full value. In this case the NEXT amplitude depends in
part on the rise time (dt) of the aggressor pulse.
10.4.1 Calculating NEXT
The backwards crosstalk coupling factor (Kb) is presented in (10.2a). It is has no
units and is used with (10.2b) when the line is short or (10.2c) when it is long to
find the actual NEXT voltage [4, 5, 8–10].
10.4
Near-End Crosstalk
113
NEXT
Saturates at
Kb × dv
Long victim
(dt < 2 tpd)
2 tpd
NEXT = 2Kb × tpd × (dv/dt)
Short victim
(dt > 2 tpd)
dv
Aggressor
dt
Voltage
Time
Figure 10.5 Width and amplitude of NEXT pulse appearing on victim depends on length of coupled region. NEXT on short lines grows in proportion to the length. It saturates at a maximum value
when the line is long.
Kb =
1
4 Ls × Cs
NEXT = 2 × Kb × tpd ×
NEXT = Kb × dV
⎛L
⎞
× ⎜ m + ( Zo × Cm )⎟
⎝ Zo
⎠
dV
dt
( when
( when
tpd = length × Ls × Cs
dt ≥ 2tpd )
dt < 2tpd )
(10.2a)
(10.2b)
(10.2c)
(10.2d)
It is evident from (10.2a) that the relationship between the trace impedance
and the coupling inductance (Lm) and capacitance (Cm) determines the NEXT
114
Understanding Crosstalk
amplitude. NEXT will always be present because, unlike with FEXT, the mutual
inductance and capacitance terms cannot cancel each other.
As shown in the Problems, when more than one trace aggresses a victim, Lm
and Cm represent the sum of the mutuals from all the aggressing traces.
Equations (10.2b) and (10.2d) show that, when the coupled length is small,
NEXT grows by increasing the length of the coupled region (tpd), by increasing
voltage swing of the aggressor (dV), or by reducing the aggressor rise time (dt).
It is apparent from (10.2c) that for long lines (defined as where the signal rise
time is less than twice the coupled length) the NEXT amplitude is only affected by
Kb and by the aggressor’s voltage swing. Therefore, NEXT is combated by lowering the coupling between traces since that lowers Kb, keeping the coupled length
short, reducing the voltage swing of the aggressor, and lengthening its rise time.
10.4.2 What Are Typical Kb Values?
Reverse crosstalk is shown in Figure 10.6 for a pair of 65Ω and 50Ω 5-mil
(0.13-mm)-wide, half-ounce thick striplines and solder mask–covered microstrips
on FR4 (εr = 4.2). The separation is given as multiples of the trace width: a separation of 3 represents 15 mils since the trace is 5 mils wide (or 0.39 mm for a 0.13-mm
trace).
It is evident that Kb (and reverse crosstalk) is greater for higher-impedance
traces, and reverse crosstalk may be larger in stripline than in microstrip.
To estimate the reverse crosstalk amplitude without using a field solver to determine Lm and Cm, Kb is found from the figure and is used in either (10.2b) or
(10.2c), depending on the length of line. The Problems present this in more detail,
0.13
0.12
0.11
0.10
0.09
65Ω
0.08
Kb
0.07
0.06
0.05 50Ω
0.04
0.03
0.02
0.01
0.00
2
4
6
8
10
Separation (trace width multiples)
Figure 10.6 Kb to an adjacent trace for 5-mil (0.13-mm)-wide half-ounce stripline (solid lines) and
solder mask–covered microstrip (dotted lines) on FR4. (Raw data from Linpar [12].)
10.5 How Closely Do Calculation and Simulation Agree?
115
but to illustrate, Kb for a 65Ω, 5-mil-wide stripline is found from Figure 10.6 to be
0.03 when the victim is a distance equal to three traces away. If the rise time of the
aggressor pulse is small compared to the electrical length of the trace, the crosstalk
voltage is found with (10.2c). For instance, NEXT will be 60 mV if the aggressor
pulse is 2V.
10.4.3 NEXT Summary
To summarize NEXT behavior:
•
The NEXT pulse has the same polarity as the aggressor pulse.
•
NEXT is possible for either stripline or microstrip traces.
•
Increasing the aggressor signal swing increases the NEXT pulse amplitude.
•
If the coupled region is long (aggressor rise time is less than twice the coupled
length, where dt < 2 tpd), the NEXT amplitude saturates at a maximum
value.
•
If the coupling occurs over a short distance (dt ≥ 2 tpd), the NEXT pulse
amplitude depends on the aggressor pulse rise time and will be smaller than
the amplitude of the long line case.
10.5 How Closely Do Calculation and Simulation Agree?
Simulation results using the SmartSpice circuit simulator [13] lossy coupled transmission line appear in Figure 10.7 for a pair of 65Ω 4-mil (0.11-mm)-wide microstrips separated by 4 mils on FR4.
We see that the simulation predicts a NEXT amplitude of 191 mV, and as
shown in the Problems, (10.2) calculates 192 mV. As expected, the NEXT pulse
has the same polarity and is as wide as the aggressor pulse. In fact, it appears as a
miniature version of the aggressor.
Two FEXT pulses are created because coupling occurs on each edge of the
aggressor pulse. The FEXT pulses are opposite polarity because (as we find in the
Problems) in this case Kf is negative. The simulation shows the FEXT pulses to be
385 mV, versus the 388 mV calculated.
The close agreement between the simulation and the equations only occurs
when the transmission line has low loss and when both ends of all the transmission lines are terminated in resistors having a value equal to Zo. Even with this
excellent correlation, the simulation and calculations are still only estimates of the
amount of crosstalk that will occur in actual hardware. This is because we usually
do not know the true shape or the actual dimensions of the fabricated traces. Nor
do we know the dielectric constant of the actual laminate or the actual solder mask
thickness. These things mean that the topology we specify in a field solver only
approximates the trace configuration that will actually be obtained in production.
116
Understanding Crosstalk
10.5.1 How Well Do 2D Field Solvers Agree?
The results shown in Figure 10.7 show that simulation and calculation give the
same results when using identical data. The correlation does not prove the field
solver has produced accurate and consistent results even if the trace topology is
known precisely. In fact, the calculation of the mutual capacitance and inductance
requires the field solver to process very small numbers, especially when the spacing
between traces is large. As shown in Table 10.1, numerical imprecision can result in
discrepancies between field solvers. To illustrate this, Table 10.1 shows the worstcase range of Kb for a 5-mil (0.13-mm)-wide, 50Ω stripline across various spacing
as reported by multiple 2D field solvers.
Table 10.1 shows that even if the trace topology is known perfectly, the simulation results obtained from different field solvers will not match exactly, especially
when the spacing is large. For instance, the field solvers agree to within 0.5% or
better for spacing of up to two times (10 mils), and the agreement is within 12% for
spacing up to six times (30 mils). It becomes significantly worse for larger distances.
10.6
Guard Traces
Crosstalk can be reduced by placing a grounded trace between a victim and aggressor to create a guard trace [7, 14–16]. When done properly, the aggressor has higher
coupling to the guard trace than to the victim, reducing the voltage induced on the
victim. Guard traces are especially effective with stripline, but they also work well
with microstrip.
One example showing the efficacy of a stripline guard trace is shown in Figure
10.8.
The top view of 4-mil (0.11-mm)-wide 65Ω stripline traces placed side by side
with a separation of 4 and 12 mils (0.11 mm and 0.32 mm) is shown. The pulse
1.5V
Aggressor
1.0V
0.5V
0.0V
NEXT
FEXT
-0.5V
0s
1 ns
2 ns
3 ns
Figure 10.7 Simulated microstrip crosstalk results. NEXT is +191 mV (versus +192 calculated). FEXT
is ±385mV (versus ±388 calculated).
10.7
Main Points
117
Table 10.1 Differences in Kb Between Multiple
2D Field Solvers for a 5-Mil (0.13-mm)-Wide, HalfOunce 50Ω Stripline on FR4
S
% Difference
1
0
2
0.5
4
3
6
12
8
45
10
170
Spacing (S) is in multiples of the trace width.
generator has an open circuit voltage of 3.3V with a 100-pS rise time. The resistors
are all 65Ω, causing a 1.65-V, 100-pS signal to be launched down each aggressor.
As Figure 10.9 shows, with a 4-mil separation NEXT is 215 mV. It falls to 55
mV when the separation is 12 mils. Inserting a 4-mil-wide guard trace uses the
same lateral space, but NEXT falls to 10 mV.
10.6.1 Connecting the Guard Trace
An improperly grounded guard trace can make crosstalk worse by coupling energy
reflecting along its length back to the victim [7, 14]. In this situation the guard trace
becomes an aggressor rather than a shield.
This can be avoided by placing a via to ground at both ends of the guard trace
and including ground vias along the length of the guard trace. These vias must have
a distance between them of less than a quarter wavelength of the highest frequency
present [14]. Shielding can be improved if the vias are placed closer than this, and
the vias should have a diameter roughly equal to the width of the trace [7].
As demonstrated in the Problems and shown in (10.3), this requirement can
be converted into a rule of thumb for FR4 if the distance (d, measured in meters)
between vias is no greater than one-tenth of the signal rise time (measured from the
10/90% points in nanoseconds) [7].
d=
tr
10
(10.3)
For instance, the vias should be separated by no more than 1/10 meter (about
4 inches) when the signal rise time is 1 ns (tr = 1 in the equation).
10.7
Main Points
•
Crosstalk is formed by inductive and capacitive coupling from one or more
aggressors to a victim.
118
Understanding Crosstalk
4 mils
215 mV
12 mils
55 mV
4 mils
4 mils
10 mV
Figure 10.8 Top view of 4-mil-wide stripline traces. NEXT falls from 215 mV at minimum spacing to
55 mV when the spacing is increased to three times the trace width. With that same overall spacing,
inserting a guard trace further reduces NEXT to 10 mV.
•
Crosstalk appearing at the end of the trace closest to the aggressor transmitter is called near-end crosstalk (NEXT).
•
Crosstalk appearing at the receiving end of the victim trace is called far-end
crosstalk (FEXT).
•
Microstrip can experience both NEXT and FEXT.
•
Stripline can experience NEXT, but FEXT will be 0 if the line has low loss.
•
The NEXT pulse has the same polarity as the aggressor pulse, and it grows
with the amplitude of the aggressor. The value saturates at a maximum value
when the line is long.
Problems
119
NEXT
0.2V
Separation = 4 mil
0.1V
Separation = 12 mil
0.0V
Guard trace
0s
1 ns
2 ns
Figure 10.9 Simulation results for Figure 10.8. NEXT falls from 215 mV when the aggressor and
victim separation is 4 mils (0.13 mm) to 55 mV when it is increased to 12 mils (0.32 mm). Inserting
a guard trace reduces NEXT to 10 mV.
•
The FEXT pulse has the same polarity as the aggressor when capacitive
coupling dominates, but has opposite polarity when inductive coupling
dominates.
•
Guard traces can be effective in reducing NEXT and FEXT, but only when
properly grounded.
Problems
Answers to these problems are available on the Artech House Web site at http://
www.artechhouse.com/static/reslib/thierauf/thierauf1.html.
10.1 A field solver has produced the following data for a pair of 8-mil-wide, halfounce striplines, separated by 10 mils:
L MATRIX [H/M]
4.447E-07
8.28E-08
8.28E-08
4.447E-09
C MATRIX [F/M]
1.088E-10
−2.025E-11
−2.025E-11
1.088E-10
Find the crosstalk coupling factors.
10.2 10.2 A field solver has produced the following data for a pair of 8-mil-wide,
half-ounce microstrips, separated by 10 mils:
120
Understanding Crosstalk
L MATRIX [H/M]
3.670E-07
5.28E-08
5.28E-08
3.670E-09
C MATRIX [F/M]
9.00E-11
−5.70E-12
−5.70E-12
9.00E-11
Find the crosstalk coupling factors.
10.3 Find FEXT for the pair of traces described in Problems 10.1 and 10.2 if the
coupled region is 40 cm (15.75 inches) in length. Assume that in each case
the signal has a rise time of 250 ps and swings from 0.8V to 1.8V.
10.4 Find NEXT for the pair of traces described in Problems 10.1 and 10.2 if the
coupled region is 40 cm (15.75 inches) in length. Assume in each case that
the signal has a rise time of 250 ps and swings from 0.8V to 1.8V.
10.5 Derive the (10.3) rule of thumb appearing in Section 10.6.1, which shows
that for guard traces on FR4 a via should be separated by no more than
one-tenth of a rise time.
10.6 The inductance and capacitance matrices for the pair of 4-mil-wide microstrips separated by 4 mils on FR4 used in the simulation appearing in
Figure 10.7 is presented here. The coupled region is 0.3 meter (11.8 inches)
in length, and the aggressor swings from 0 to 1.68V in 0.1 ns. Calculate
NEXT and FEXT.
L MATRIX [H/M]
4.009E-07
9.662E-08
9.662E-08
4.009E-09
C MATRIX [F/M]
1.029E-10
−2.233E-11
−2.233E-11
1.029E-10
10.7 Use Figure 10.6 to find Kb for a pair of 5-mil-wide half-ounce 50Ω striplines on FR4 that are separated by 10 mils.
10.8 The inductance and capacitance portion of an RLGC SPICE model appears
below. What NEXT will appear on the center trace if the two flanking
traces are aggressors, when the aggressors each launch 1.65V pulses with a
0.1-ns rise time? Assume that the coupled length is 0.3 meter (11.8 inches)
and that the model data is presented in meters.
+ Lo = 4.380E-07
+ 1.124E-07 4.327E-07
+ 3.405E-08 1.124E-07 4.380E-07
+ Co = 1.143E-10
+ −2.935E-11 1.233E-10
+ −1.356E-12 −2.935E-11 1.143E-10
Problems
121
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
Young, B., Digital Signal Integrity, Upper Saddle River, NJ: Prentice-Hall, 2001.
DeFalco, J. A., “Reflections and Crosstalk in Logic Interconnections,” IEEE Spectrum, July
1970, pp. 44–50.
Paul, C. R., Analysis of Multiconductor Transmission Lines, New York: John Wiley &
Sons, 1994.
Hart, B. L., Digital Signal Transmission Line Circuit Technology, Oxford, U.K.: Van Nostrand Reinhold, 1988.
Rosenstark, S., Transmission Lines in Computer Engineering, New York: McGraw-Hill,
1994.
Catt, I., “Crosstalk (Noise) in Digital Systems,” IEEE Transactions on Electronic Computers, Vol. EC-16, No. 6, December 1967, pp. 743–763.
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Feller, A., et al., “Crosstalk and Reflections in High-Speed Digital Systems,” Proceedings of
the Fall Joint Computer Conference, Washington, D.C., December 1965, pp. 511–515.
Sengupta, M., et al., “Crosstalk Driven Routing Advice,” Proceedings of the 44th Electronic Components and Technology Conference, Washington, D.C., May 1994, pp.
687–694.
Shi, W., “Evaluation of Closed-Form Crosstalk Models of Coupled Transmission Lines,”
IEEE Transactions on Advanced Computing, Vol. 22, No. 2, May 1999.
Polar Si9000 Field Solver, version V9.01.00, Polar Instruments, Inc. Beaverton, OR.
Djordjevic, A. R., et al., LINPAR for Windows, Norwood, MA: Artech House, 1999.
SmartSpice, Simucad Design Automation, Version V3.10.8.R, Santa Clara, CA.
Ladd, D. N., et al., “SPICE Simulation Used to Characterize the Cross-Talk Reduction Effect of Additional Tracks Grounded with Vias on Printed Circuit Boards,” IEEE Transactions on Circuits and Systems II: Analog and Digital Signal Processing, Vol. 39, No. 6, June
1992.
Bakoglu, H. B., Circuits, Interconnections, and Packaging for VLSI, Reading, MA: Addison-Wesley, 1990.
Suntives, A., et al., “Using Via Fences for Crosstalk Reduction in PCB Circuits,” 2006
IEEE International Symposium on Electromagnetic Compatibility, Portland, OR, Vol. 1,
August 1–18, 2006, pp. 34–37.
C H A P T E R 11
Understanding Signal Reflections
11.1
Introduction
The previous chapters have assumed that the transmission line has the same impedance everywhere and that the load impedance is the same as the impedance of the
line. Although this is usually the goal, especially when signaling at high speeds, it is
often not achieved in practice.
Reflected energy waves (signal reflections) created when the transmission line
impedance does not equal the receiver impedance affect the voltage and shape of
the waveform appearing at the receiver terminals. The waveform becomes even
more distorted when the reflections created by a mismatch in the impedance between the driver and the transmission line arrive back at the load. The change in
the waveform can be quite dramatic, as illustrated in Figure 11.1.
Figure 11.1 shows the voltage appearing at a receiver having a high input impedance when a 3.3-V square wave is launched down a transmission line whose
impedance does not match either the driver or the receiver. The distorted wave
shape is obvious, but the voltage levels that the pulse reaches (+5V and −2V) and
the length of time that it is at those levels are also important. As discussed in Chapter 1, these voltages and times may be great enough to damage the ASIC I/O cell.
In this chapter we will examine the behavior when the source and load impedances do not equal the transmission line impedance, and see how it is possible for
waveforms such as those appearing in Figure 11.1 to be created. Chapter 12 expands on this and presents strategies on how to correct these mismatches and how
to restore the waveform and maintain proper voltage levels.
11.2 How Are Reflections Created?
A pulse travels down a lossless transmission line in the setup appearing in Figure
11.2.
The launched voltage is observed at Vne and a current pulse of the same shape
is simultaneously launched. Together these two pulses represent the wave of energy
(the incident wave) traveling down the line from the generator resistance (Rg) to the
123
124
Understanding Signal Reflections
Receiver input voltage
6V
4V
2V
0V
−2V
Time
Figure 11.1 A launched 3.3-V square wave measured at the receiver when neither the transmitting
nor the receiving impedances equal the transmission line impedance.
(+) Direction
(−) Direction
Rg
Vfe
Vne
Rl
Figure 11.2 Transmission line connected to a pulse generator of resistance Rg and a load resistance
Rl. Waves move in the positive direction when flowing left to right. The negative direction is right
to left.
load resistance (Rl). Waves traveling from left to right are defined to be moving in
the (+) direction, while right-to-left movement is in the (−) direction.
As first discussed in Chapter 6, and repeated in (11.1), the generator launches
incident voltage (Vi) and incident current (Ii) waves in a ratio determined by the
transmission lines characteristic impedance (Zo).
Zo =
Vi
Ii
(11.1)
This ratio is maintained everywhere along the line unless there is a change in
the impedance. However, difficulties arise when the load resistance does not match
the line impedance, as demonstrated in Figure 11.3.
In Figure 11.3(a) an incident voltage pulse of 1V is launched down a 50Ω
transmission line. From (11.1) an incident current pulse of 20 mA accompanies the
11.2
How Are Reflections Created?
125
50Ω
20 mA
20 mA
1V
50Ω
V=I×R
= 20 mA × 50Ω
= 1V
R l = 50Ω
2V
(a)
50Ω
20 mA
20 mA
1V
50Ω
V=I×R
= 20 mA × 0Ω
=0
2V
(b)
50Ω
20 mA
0 mA
1V
50Ω
I=
V 1V
=
=0
R
∞
2V
(c)
Figure 11.3 Launched voltage and current remains in the proper ratio (a) when the load resistance
equals the transmission line resistance, but not when the load is (b) a short circuit or (c) an open
circuit.
voltage pulse. Equation (11.1) is still satisfied when the pulses reach the 50Ω load
resistor, but this is not the case when the load has any other value.
For instance, changing the load to a short circuit as is shown in Figure 11.3(b)
requires the load voltage to be zero. However, the energy represented by the incident voltage and current pulses cannot simply vanish when the waves reach the
short-circuited load. A similar problem shown in Figure 11.3(c) occurs when the
load is infinite, representing an open circuit. There the current must become 0 at
the load, even though 20 mA has been launched down the line.
The solution to these dilemmas centers on conservation of energy. The incident
energy wave launched by the generator is perfectly absorbed by the load when Rl
matches the transmission line impedance, but this is not so for any other value of
load resistance. For instance, the launched energy cannot be absorbed by the open
circuit, and because the line is lossless the energy is not dissipated there. Instead,
when the energy wave encounters the infinite impedance of the open circuit, it rebounds and travels back up the line toward the generator. The rebound is a reflection, and wave theory is used to explain its behavior [1–4].
126
Understanding Signal Reflections
11.2.1 A Physical Analogy for Transmission Line Reflections
The reflection can be pictured by a physical analogy of a pulse traveling along a
rope. For this analogy the transmission line in Figure 11.3(c) has been replaced with
a taught rope. The far end is securely attached to a massive wall, representing the
infinitely high impedance of an open circuit.
By holding and then abruptly shaking the near end, an incident pulse is launched
that travels toward the far end. From elementary physics (for example, see [5]),
when the incident pulse strikes the immovable wall, the pulse reflects, sending a
smaller pulse of the same polarity back up the rope to the source.
In a similar way, the incident voltage and current pulses traveling in the (+)
direction, down the transmission line, reflect from the open circuit. These reflected
waves travel as freely along the line as do the incident waves and are called the
reflected voltage (Vr) and the reflected current (Ir) waves. They travel at the same
velocity as the incident waves, but in the (−) (that is, the opposite) direction.
11.3
The Reflection Coefficient
The percentage of the incident voltage wave that is reflected from an impedance
discontinuity is called the voltage reflection coefficient, ρ. The current reflection
coefficient is ρi. These are used in (11.2), where Vi and Ii are the incident voltages
and current waves, and Vr and Ir are the values of the reflected waves.
Vr = Vi × ρ
(11.2a)
Ir = Ii × ρi
(11.2b)
These relationships apply equally to the receiving and transmitting ends of the
transmission line.
Although the voltages and currents have separate reflection coefficients, when
signal integrity engineers refer to the reflection coefficient, they generally mean the
voltage reflection coefficient. In this book whenever ρ is used it will represent the
voltage reflection coefficient, and ρi will be used in those instances when it is necessary to indicate the current reflection coefficient.
11.3.1 What Determines the Reflection Coefficient Value?
The reflection coefficient for a transmission line with low loss is given in (11.3). Its
derivation relates the ratio of the incident and reflected waves.
ρ=
R − Zo
R + Zo
(11.3)
11.4 How Do the Reflected Waves Combine with the Incident Waves?
127
The transmission line impedance is Zo, and R is either Rl (the resistance of the
load resistor), or Rg (the generator resistance). The reflection coefficient at the load
is ρl; it is called ρg at the generator.
For example, ρl is −0.25 when a 25Ω load is connected to a 50Ω transmission
line, and it is 0 when the load equals 50Ω. The reflection coefficient at the load
becomes +0.25 when the load resistance is 83Ω. The Problems have further examples, but these show that ρ is positive when the load resistance is greater than the
transmission line impedance. When it is less, ρ is negative. It becomes zero when R
is the same as the transmission line impedance. It can be shown [6, 7] that the current reflection coefficient has the same value but the opposite sign from the voltage
reflection coefficient. These results are summarized in Table 11.1.
Equation (11.3) can still be used if the loads are complex (such as capacitors
or inductors rather than resistors), or when the transmission line impedance is
complex (as is the case when the line is lossy). In those cases ρ is replaced with Γ
(uppercase gamma), and in Figure 11.3 R becomes the load and generator impedances Zl or Zg.
11.4 How Do the Reflected Waves Combine with the Incident Waves?
From wave theory [1, 3] we know that the total voltage at the load is the sum of
the incident and reflected waves (the waves superimpose). As shown algebraically
in (11.4), this combined wave superimposes with any portions of the wave still
present.
Vt = Vi + Vr + Vp
(11.4a)
It = Ii + I r + I p
(11.4b)
The total voltage and currents are Vt and It, Vi and Ii are the incident pulses,
and Vr and Ir are the reflected voltage and current pulses. Vp and Ip are the voltage
and currents still present.
11.4.1 Difference Between the Response of a Pulse and a Step
The third term in (11.4) hints at the interaction between the pulse width and the
electrical length of the line. If the pulse width is very wide, the reflected signal
will arrive in time to combine with it. However, if the pulse is narrow, it will have
changed state before the reflection arrives.
Table 11.1
Reflection Coefficient Values
R
ρ
ρi
Notes
0
−1
+1
R is much less than Zo (short circuit)
Zo
0
0
R equals transmission line Zo (matched circuit)
∞
+1
−1
R is much larger than Zo (open circuit)
128
Understanding Signal Reflections
By examining Figure 11.4(a), we see that when the pulse (Vp) is narrow enough
to return to zero before reflected pulse arrives, the total near-end voltage (Vt) is the
sum of the incident and reflected pulses.
However, as shown in Figure 11.4(b), if the pulse is wide enough to still be
present when the reflection arrives, the incident and reflected pulses add during the
period when the pulses overlap, combining to change the shape of the waveform.
Although in Figure 11.4 the incident and reflected waves are positive going, they
can be negative going.
11.5 What Is the Behavior When There Are Multiple Reflections?
Commonly in digital systems an impedance mismatch is present at both ends of the
transmission line. This causes reflections to rebound from both ends, which can
significantly alter the wave shape and signal amplitude at both ends of the line. A
demonstration of this setup appears in Figure 11.5.
Vp
(Vi + Vr )
Vt
2td
Vp
(Vi + Vr )
Vt
Voltage
Time
Figure 11.4 The total voltage (Vt) (a) when the launched (Vp) pulse is narrow and (b) when it is
wide enough to combine with the reflection.
11.5 What Is the Behavior When There Are Multiple Reflections?
129
50Ω
Vne
Rg = 10Ω
Vfe
2 ns
2V
200 ps
Figure 11.5 Circuit setup demonstrating reflections when the generator and load resistances do
not match the transmission line impedance.
Figure 11.5 shows a 2-V, 10Ω driver connected to a 60Ω, 2-ns-long lossless
transmission line. The far end of the line is open-circuited, representing the input
of a high-impedance receiver. The driver launches a single pulse that has a width
of 200 ps, which is shorter than the electrical length of the line, and very fast rise
and fall times. A circuit simulation showing the behavior of this circuit appears in
Figure 11.6.
Figure 11.6(a) shows the load voltage, Vfe. The voltage’s near end, Vne, is
shown in Figure 11.6(b). Multiple reflections are present and it is apparent that the
voltage at the near and far ends alternate between a (+) going and (−) going wave.
The voltage at the load caused by the first reflection actually exceeds the driver’s
2-V supply voltage.
Hand calculations can recreate the simulation results, but because there are
multiple reflections, keeping track of the various sums becomes cumbersome. A
4
3.34V
V fe
2
1.5V
0
-0.9V
−2
−2.24V
(a)
4
Vne
2 1.67V
0.55V
0.25V
0
−0.37V
−2
0.0s
2.0 ns
4.0 ns
6.0 ns
8.0 ns 10.0 ns 12.0 ns 14.0 ns
(b)
Figure 11.6
tor (Vne).
Simulation results for Figure 11.5. Voltages (a) at the load (Vfe) and (b) at the genera-
130
Understanding Signal Reflections
reflection chart [1, 6, 7] will be used manage the sums and to show the timing and
amplitude of the reflections.
We begin by using the voltage divider principle to determine that the driver
launches 1.67V down the line. This is the incident voltage, Vi.
From (11.3) we find the reflection coefficient at the driver end of the line is
Rg − Zo 10 − 50
ρg =
=
= −0.67 , and from Table 11.1, ρl is +1.
Rg + Zo 10 + 50
We are now ready to construct the reflection chart shown in Figure 11.7.
The voltages present at the generator end of the line are recorded on the left
side of the chart, and the load end voltages are recorded on the right side. Time
increases vertically downward with the zigzagging lines representing the reflection
bouncing between the two ends. Creating a reflection chart for the voltage wave is
described here, but the method for the current chart is the same.
The chart is filled in as follows:
1. At t = 0 the 1.67V incident wave found with the voltage divider principle
is launched.
Generator
end
Load
end
t=0
1.67V launched
Vi
=1
.67
V
t = 2 ns
r l = +1
Vt = (1.67) + (1.67) = 3.34V
1
+
ρ=
t = 4 ns
rg = −0.67
Vr
Vt = (1.67) + (−1.12) = 0.55V
ρ=
Vr
V
.67
=1
−0
.6
7
=−
1.1
2
t = 6 ns
Vt = (−1.12) + (−1.12) = −2.24V
ρ=
.67
−0
Vr
t = 8 ns
Vt = (−1.12) + (0.75) = −0.37V
2
1.1
=−
ρ=
Vr
=0
+0
.45
.75
t = 10 ns
Figure 11.7 Reflection chart illustrating results from Figure 11.6. The load is separated from the
generator by an electrical distance of 2 ns. Generator voltages appear on the left and load voltages
appear on the right.
11.5 What Is the Behavior When There Are Multiple Reflections?
131
2. At t = 200 ps the output of the pulse generator returns to 0V.
3. The pulse arrives at the load at t = 2 ns. Since ρl = +1, a 200-pS-wide reflection of 1.67V is created. From (11.4) the load voltage is the sum of the incident and reflected voltages, noting that the load was previously 0V. This
makes the load voltage 3.34V.
4. The +1.67V reflection (and not the 3.34V load voltage) travels back down
the line and reaches the generator at t = 4 ns. Since ρg = 0.67, a 200-pS reflection of −1.12V is created from the +1.67V reflection impinging on the
generator (the output of which is 0V). From (11.4) the generator voltage
now becomes 0.55V.
5. The reflected −1.12V, 200-pS-wide pulse travels back up the line and at t
= 6 ns arrives at the load where it sums with the −1.12V reflection created
by the open circuit. Because the load has returned to zero by the time this
reflection arrives, a total voltage of −2.24V is created.
6. The −1.12V, 200-pS-wide reflection travels down the line toward the generator, where it arrives at t = 8 ns.
7. When the −1.12V reflection reaches the generator, a +0.75V reflection is
produced which reaches the load at t = 10 ns. As the chart shows, this is
45% of the initial launched voltage of 1.67V. In fact, this is the product of
the percentages of all the proceeding reflections (1 × −0.67 × 0.67). This
observation can be used to find the voltage after any number of reflections
without resorting to constructing a reflection chart.
8. This process continues with the reflection becoming smaller and changing
signs each time it rebounds from the generator impedance.
From the reflection chart it is apparent how it is the energy from the first
reflection that makes its way up and down the transmission line. Because each
subsequent reflection is multiplied by ρ (whose absolute value is never greater than
1), this means that, ignoring the effects of crosstalk, in a single pulse system the
subsequent reflections will never have a magnitude worse than the first reflection.
This knowledge is useful when setting up transmission line simulations and when
making laboratory measurements.
The chart suggests a difference between pulse and step excitation. If the signal
is a step longer than twice the round trip of the transmission line, the reflections
would add to the launched voltage to create the composite voltage, Vne. For instance, the near-end voltage at step 4 in the reflection chart (Figure 11.7) would
have become 2.22V (the sum of the reflections plus the 1.67V still present from the
generator) if the launched pulse was a step.
Reflection charts are no longer a practical design tool because even complex
networks can easily be modeled with modern circuit simulators. However, because
they show the behavior and characteristics of the reflected energy, they are a good
learning tool that the signal integrity engineer should master.
132
Understanding Signal Reflections
11.5.1 What Is the Behavior When There Are Many Pulses?
The load voltage when a stream of pulses (rather than a single pulse) is launched
down the line appears in Figure 11.8. The setup is identical to that used to create
Figure 11.6, but in this case the 200-ps pulses repeat every 500 ps.
Assuming that the receiver switches at 1.65V (corresponding to
Vdd
), the re2
ceiver will properly detect the first eight pulses (Vfe = 3.34V), but the amplitude of
the following eight pulses are too low to register as logic 1s. The next group will be
properly detected, but the last group has marginal amplitude.
It is apparent from this how the data pattern influences the received signal,
even when the transmission line has no losses. Proper termination can greatly mitigate these effects, but as part of the analysis process the signal integrity engineer
should determine the set of data patterns that make the interconnect behave in the
worst possible way. Once identified, any that can legitimately occur in the system
should be included for use during design verification testing measurements on actual hardware.
Reactive Discontinuities
The loads connected to transmission lines in actual systems always include some
amount of capacitance and inductance. For instance, the leads of an integrated circuit or discrete component are often inductive, and a via can appear capacitive. In
Chapter 14 we will see how bends add excess capacitance and inductance to a trace.
The way in which these and other discontinuities that can appear along a transmission line are sketched in Figure 11.9.
4
3.34V
2.6V
V fe
2
1.6V
1.1V
0
−2
4
2.2V
1.67V
2
Vne
11.6
1.9V
2.1V
0
−2
0.0s
2.0 ns
4.0 ns
6.0 ns
8.0 ns 10.0 ns 12.0 ns 14.0 ns
Figure 11.8 Effects of multiple pulses on the circuit of Figure 11.5. Vfe and Vne are significantly different than the otherwise identical single pulse results illustrated in Figure 11.6.
11.6
Reactive Discontinuities
133
Vne
Zo = 50Ω
Zo = 50Ω
Vfe
50Ω
Shunt capacitor
1V
Vne
Zo = 50Ω
Zo = 50Ω
Vfe
50Ω
Series capacitor
1V
Vne
Zo = 50Ω
Zo = 50Ω
Vfe
50Ω
1V
Shunt inductor
Vne
Zo = 50Ω
Zo = 50Ω
Vfe
1V
50Ω
Serries inductor
Figure 11.9 Types of reactive discontinuities.
From circuit theory we know that a capacitor initially acts as a short circuit,
and how the voltage exponentially rises as it charges. Once fully charged, it appears as an open circuit. From an impedance standpoint the capacitor initially has
a low impedance that exponentially increases. If given enough time, the impedance
becomes infinite.
Similarly, an inductor initially acts as an open circuit and it charges exponentially to become a short circuit. Its impedance is initially infinite and falls exponentially to zero.
Table 11.1 showed that a short circuit creates a negative reflection, while an
open circuit creates a positive one. The result is that a shunt capacitor appears as
a sharp negative going reflection that exponentially returns to the base line voltage
as the capacitor charges and the impedance increases. A series inductor initially
134
Understanding Signal Reflections
appears as a positive reflection that exponentially returns to the baseline as its impedance falls. These behaviors are evident in Figure 11.10(a, d).
The charging behavior of the series capacitor configuration is apparent in Figure 11.10(b), and the shunt inductor in Figure 11.10(c). Initially the capacitor is
a low impedance and its impedance increases as it is charged by the launched current. The inductor is a high impedance, but as the inductor charges, its impedance
decreases, shunting increasingly more of the signal energy to ground.
11.7
Main Points
•
The energy transmitted down a transmission line consists of voltage and current waves.
•
Conservation of energy requires that a reflection wave be created from the
incident wave when a change in impedance is encountered.
•
The reflection coefficient (ρ) indicates how much of the incident wave is reflected from a discontinuity.
•
An open circuit creates a positive voltage reflection and a negative current
reflection.
1.00V
1.00V
0.75V
0.75V
Series capacitor
Shunt capacitor
0.50V
0.50V
0.25V
0.25V
0.00V
0.0s
1.0 ns
(a)
0.0s
1.0 ns
(b)
1.00V
0.75V
0.00V
1.00V
0.75V
Shunt inductor
Series inductor
0.50V
0.50V
0.25V
0.25V
0.00V
0.0s
1.0 ns
(c)
0.0s
1.0 ns
(d)
Figure 11.10 (a–d) Waveforms created by discontinuities as measured at Vne for a step voltage.
0.00V
Problems
135
•
A short circuit creates a negative voltage reflection and a positive current
reflection.
•
At the discontinuity the incident and reflected waves add to create a composite wave.
•
Multiple reflections can be simultaneously present on a line and can combine
in complex ways.
•
Reflections from earlier pulses can combine with later pulses to change their
shape and amplitude.
Problems
Answers to these problems are available on the Artech House Web site at http://
www.artechhouse.com/static/reslib/thierauf/thierauf1.html.
11.1 Find the voltage reflection coefficient at the load when Rg = 10Ω, Rl =
100Ω, and the transmission line has an impedance of 75Ω.
11.2 Using the same information from in Problem 11.1, find the voltage reflection coefficient at the generator.
11.3 Using the information from the previous problems, show how the incident
wave does not satisfy Ohm’s law when it reaches the load. Then show how
the reflection determined by ρ causes the network to satisfy Ohm’s law.
11.4 What value of voltage and current will be reflected back from the load in
Figure 11.2 when the load is an open circuit? Assume that the pulse generator has an open circuit voltage of 2.2V and an impedance Rg of 25Ω and
that the transmission line impedance is 50Ω.
11.5 Assuming in Figure 11.2 that Vne is a 1-V pulse when the pulse generator is
not connected to a load, determine the Vfe when Rg = 50Ω, Zo = 50Ω, and
Rl = 100Ω.
11.6 Assuming in Figure 11.2 that the transmission line has an electrical length
of 2 ns and that Vne is a 200-ps-wide, 1-V pulse when the pulse generator
is not connected to a load, create a reflection chart for Vfe when Rg = 25Ω,
Zo = 50Ω, Rl = 100Ω.
11.7 Assume in Figure 11.2 that the transmission line is lossless and has an electrical length of 2 ns. If the pulse generator creates a 200-ps-wide, 1-V pulse
when it is not connected to a load, describe the behavior at the near and far
ends when Zo = 50Ω and is Rl = ∞ (an open circuit). Assume that Rg = 150Ω
when the pulse is first launched, but that Rg becomes an open circuit before
the first reflection arrives back at the source.
References
[1]
[2]
[3]
Brown, R. G., et al., Lines, Waves and Antennas, New York: John Wiley & Sons, 1973.
Johnson, W. C., Transmission Lines and Networks, New York: McGraw-Hill, 1950.
Skilling, H. H., Transient Electric Currents, New York: McGraw-Hill, 1952.
136
Understanding Signal Reflections
[4]
[5]
[6]
[7]
Matick, R. E., Transmission Lines for Digital and Communication Networks, New York:
IEEE Press, 1969.
Serway, R. A., Physics for Scientists and Engineers, 4th ed., Philadelphia, PA: Sanders College Publishing, 1996.
Rosenstark, S., Transmission Lines in Computer Engineering, New York: McGraw-Hill,
1994.
Hart, B. L., Digital Signal Transmission Line Circuit Technology, Berkshire, U.K.: Van
Nostrand Reinhold, 1988
C H A P T E R 12
Termination Strategies
12.1
Introduction
In the previous chapters we implied that the load and transmission line impedances must match for proper electrical operation of an interconnect. However, the
actual goal is to ensure that the received signal crosses the receiver threshold with
adequate timing margin and, at least for clocks, that the signal transitions smoothly.
The signal integrity engineer accomplishes this by arraigning the wiring topology and selecting the termination scheme and I/O drivers so that reflections do
not cause multiple pulses, glitches in the waveform, or unacceptable voltage levels.
Terminations can be placed at the load end (far-end termination) or at the driving
(source termination) end. The choice depends on available timing margins, the
wiring topology, the number of loads on a net and their location compared to one
another, and power dissipation.
12.2 Source Series Termination
Source series termination uses a resistance to connect the driver (the source) to the
transmission line. This resistance raises the impedance of the transmitter (TX in Figure 12.1) and is either a discrete resistor placed on the circuit board (shown as Rs),
or it is an integral part of the integrated circuit I/O cell. In some integrated circuits
the built-in series impedance can be selected (programmed) from a range of values.
12.2.1 Selecting the Resistance Value
Until now, we have tacitly assumed that the driver impedance has the same value
as the transmission line impedance. However, in some circumstances it can be advantageous for the driver impedance not to equal the transmission line impedance.
Three cases are shown in Figure 12.2 for the point-to-point configuration
shown in Figure 12.1. The driver impedance is very low so that series resistor Rs
sets the total impedance value.
137
138
Termination Strategies
Vdd = 1.8V
Zo
RS
TX
RX
V ne
Vtt = 0.9V
Figure 12.1 Source series termination. Resistor Rs may be incorporated in the driver or, as shown,
may be a discrete resistor on the circuit board. Vtt sets the receiver switch point.
2.5V
Rs = 25Ω
Rs = 50Ω
2.0V
1.5V
1.0V
Receiver
switch point
Rs = 75Ω
0.5V
0.0V
−0.5V
−1.0V
15 ns
20 ns
25 ns
Figure 12.2 Response at the load when a 1.8-V driver is connected to a 50Ω transmission line with
Rs equal to 25Ω (broken curve), 50Ω (solid), and 75Ω (dotted).
From Figure 12.2 we see that by making the series resistance (Rs) equal to the
50Ω impedance of the transmission line, the received signal switches cleanly from
0V to 1.8V (the value of Vdd, the voltage powering the driver).
By making Rs smaller than Zo (in this case, 25Ω), the driver launches more
current and a higher voltage down the transmission line, and the waveform crosses
the receiver switch point sooner. However, the disadvantage is that any reflections
arriving back at the driver from the load will be rereflected. In fact, the generator
reflection coefficient (described in Chapter 11) will be negative, meaning that once
the reflection arrives back at the load, it will subtract from the original wave. If
large enough, this can cause the signal to fall (ring back) below the point where the
receiver switches. In this case the switch point is set by Vtt to be half of Vdd, and
since the signal only dips to 1.5V, the receiver does not falsely trigger. However, the
signal exceeds Vdd (1.8V) and transitions significantly below ground. The signal
integrity engineer would need to confirm that these voltages do not exceed safe
operating limits for the integrated circuit technology.
Making Rs higher than Zo (75Ω in Figure 12.2) causes less current and a lower
voltage to be launched, and although the reflections do not exceed Vdd or go below
ground, the lower current has caused the switch point to be crossed later.
12.3 Parallel and Thevenin Termination
139
If this timing degradation is acceptable, launching less current can be an advantage in low-power applications. However, the signal rise time will suffer if Rs
is made too large, and care should be taken to ensure that the rise time is not degraded so much that the signal slowly transitions through the receiver switch point.
Such slow transitions make the receiver susceptible to noise induced false triggering
and can cause the receiver to switch multiple times before the signal is large enough
to be properly detected. The resulting unpredictable timing and the creation of errant pulses (chattering) is never acceptable for clocks or strobe signals since it can
lead to double clocking. Although the logic design may tolerate this type of behavior on signal lines, chattering is still undesirable as it can cause the receiver to draw
excessive current and can generate power supply noise that may cause problems
elsewhere in the system (including increased RF emissions).
12.2.2 Effect of Driver Resistance
The output transistors in practical CMOS I/O drivers always have some on resistance (Rds_on) that adds to the resistor Rs placed on the circuit board. Because Rds_on
is nonlinear, its value depends on the current being launched by the driver. This is
why it is poor practice to use a simple resistor to model I/O behavior rather than an
actual circuit model of the driver.
Additionally, in CMOS I/O drivers Rds_on generally increases as the power supply is reduced and as temperature increases. Circuit simulation shows these effects
and should be used to determine the value of Rs for proper operation across the
range of transmission line impedances, power supply voltages, and temperatures
expected in the system. The tolerance and the temperature coefficient of the resistor
type used for Rs should be included in the model.
12.2.3 Power Savings
If we assume that the connection from the resistor to the driver is electrically short,
Rs simply adds to Rds_on, limiting the amount of current that the driver can launch.
Since the current is less, the power dissipated by the driver is reduced. A second
benefit of adding Rs is that the total power is shared between the driver and the resistor mounted on the circuit board. In effect, by adding Rs, the signal integrity engineer has moved some of the power dissipation from the ASIC to the circuit board.
In some instances the increased number of circuit board components is worth the
beneficial effect of helping to lower the ASIC junction temperature.
12.3 Parallel and Thevenin Termination
Parallel termination is placed at the load end of the transmission line. The two principal variants are shown in Figure 12.3.
The driver TX can be either a push-pull or an open drain type driver. The pushpull driver sources current when launching a logic high and sinks current to pull the
line to a logic low. The open drain (or open collector when the I/O cell is made with
bipolar logic) driver only sinks current to pull the line low. In this case the parallel
termination supplies the current required to pull the line up to a logic high.
140
Termination Strategies
Vdd
Zo
V ne
R1
Vfe
TX
RX
R2
Vtt
(a)
Vdd
Zo
Vfe
TX
RX
Rterm
Vterm
Vtt
(b)
Figure 12.3 (a) Principal types of parallel termination. R1 and R2 form a voltage divider from Vdd
with an impedance equal to Zo. (b) The Thevenin equivalent.
In the general parallel termination scheme, two resistors form a voltage divider
from a power supply (usually the supply powering the driver, Vdd, but this is not
always the case). Ideally the supply has a low impedance and there is an AC short
circuit from Vdd to ground. This places resistors R1 and R2 in parallel, but because
in actual hardware the power supply has parasitic series inductance, it has a large
impedance at high frequencies. To compensate for this, the AC short is created by
adding decoupling capacitors between Vdd and ground close to R1 and R2. This
makes the impedance of the termination (Rterm) the parallel combination of R1 and
R2, as shown in (12.1).
Rterm =
R1 × R2
R1 + R2
(12.1)
The termination voltage (Vterm) is determined by the voltage divider principle,
shown in (12.2).
Vterm = Vdd ×
R2
R1 + R2
(12.2)
12.3 Parallel and Thevenin Termination
141
Resistors R1 and R2 can be found in terms of Vdd and Vterm by simultaneously
solving these two equations, producing (12.3a) and (12.3b).
R1 =
Vdd × Rterm
Vterm
(12.3a)
R2 =
R1 × Vterm
Vdd − Vterm
(12.3b)
The Problems show how to use these equations, but as an example, to obtain
Rterm of 50Ω and 1.65V for Vterm when Vdd is 5V, we first use (12.3a) and find
152Ω for R1. This value is used in (12.3b) to find that R2 is 75Ω. Once these values
are known, (12.1) and (12.2) can be used to determine how changes in resistance
due to temperature or component tolerance affects Rterm and how changes in Vdd
affects Vterm.
Resistors R1 and R2 can be separate, discrete resistors or a resistor network
component. The matching between resistors is superior with the networks, but
crosstalk and ground bounce can be worse than when using discrete devices. In
general, surface mount devices have lower parasitic inductance and offer better
high-frequency response than through-hole devices, with BGA devices giving the
best response [1].
However, when using resistor networks, it is important to verify by simulation
(and later during DVT) that crosstalk and simultaneous switching effects do not
cause unacceptable amounts of noise for those terminators that share a common
power and ground connection within the device.
12.3.1 Selecting Vtt and Vterm
Generally in CMOS systems, the driver switches “rail-to-rail”: the output swings
between Vdd and ground. In these systems the receiver switch point is set by a termiV
nation voltage (Vtt in Figure 12.3) to be half the value of Vdd ( dd ), since this cre2
ates symmetrical low and high noise margins. For instance, 120Ω resistors would
V
be used for R1 and R2 to terminate a 60Ω transmission line at dd . Assuming Vdd
2
is 2.5V, this gives a 60Ω termination centered at Vtt = 1.25V.
However, depending on the switching technology, half of Vdd is not always the
most appropriate value for Vtt. For instance, in TTL signaling (originally implemented with bipolar technology but now largely done with CMOS or BiCMOS), a
logic low is sensed with a voltage close to ground and a high with a voltage nearly
equal to half of Vdd. As explored in the Problems, in this situation the termination
voltage is shifted to be midway between the two logic values.
The voltage for Vterm must be chosen carefully when working with threestate buses (that is, a multidrop bus where the transmitters can be placed in a
142
Termination Strategies
high-impedance state). In some power-sensitive applications, all of the drivers are
turned off when the bus is not actively driving data. As shown in the Problems, this
greatly reduces the power dissipated by the termination compared to the situation
where the bus is statically held in a state (often low) for extended periods. However, with all the transmitters turned off, resistors R1 and R2 force the bus voltage
to become equal to Vterm. If this is close to Vtt, the result may be receiver chattering
and the creation of power supply or RF noise, along with an increase in receiver
power dissipation. These effects are especially pronounced on bidirectional buses
where each node on the bus has a three-state driver and receiver. In this situation
the termination voltage should be selected that is guaranteed to exceed the receiver
switch point under all environmental and worst-case tolerance conditions.
12.3.2 Thevenin Termination
From circuit theory the two resistors and the Vdd power supply in Figure 12.3 can
be replaced with a Thevenin equivalent resistance in series with a Thevenin voltage
supply [shown in Figure 12.3(b) as Rterm and Vterm]. This Thevenin termination is
often used to terminate signals driven to memory devices, such as double data rate
(DDR) memories.
For instance, a 60Ω Thevenin termination centered at 1.25V is formed with a
single 60Ω resistor (Rterm) connected to a 1.25-V power supply (Vterm). This approach uses half the number of resistors and generally allows the single resistor
to be placed closer to the point of termination than is possible with two resistors.
When done properly, this improves signal quality and, when many lines are terminated, can be lower in cost. However, the creation of Vterm must be done with
care, and in small systems it may cost more to properly create Vterm than is saved
by eliminating half of the termination resistors.
It is generally inappropriate to use a simple three-terminal voltage regulator for
the Vterm supply. Those regulators usually only source current, but the termination
requires that it both source and sinks current. Instead, regulators especially designed for termination applications should be used, and the manufacturer’s guidelines concerning placement and value of decoupling capacitors must be closely
followed.
High amounts of rapidly changing current flow when many signals are terminated, and to keep the termination voltage noise-free, the circuit board layout must
provide a low-inductance path from the terminator resistor to Vterm.
Narrow traces are too inductive to connect the regulator to the terminators.
For this reason in large, high-performance systems, a voltage plane (or a portion
of a plane) is sometimes dedicated to supplying Vterm, much as a plane is used to
supply Vdd or ground. Besides providing a low-inductance connection from Rterm
to Vterm, a plane frees up routing channels that would otherwise be consumed by
wide traces feeding Vterm to various locations.
As explored in the Problems, Thevenin termination can dissipate less system
power than parallel termination, especially when a three-state bus is intentionally
placed in the high-impedance state to save power.
12.4
Diode Termination
143
12.3.3 Connecting Rterm Directly to Vdd or Ground
A further refinement on the Thevenin termination is possible by setting Vterm to
either 0V or equal to Vdd. As shown in Figure 12.4, this is equivalent to connecting
the resistor directly to ground or to the power supply.
This scheme can be helpful in reducing power dissipation if the signal is predominately in one state. For example, by connecting the termination to ground,
a driver sending a pulse that is most often at ground potential will only dissipate
power during those infrequent times when the signal voltage pulses above ground.
12.4
Diode Termination
The scheme shown in Figure 12.5 shows that instead of preventing reflections by
impedance matching the load to the line, reflections can be clipped by replacing
resistors R1 and R2 with diodes. Signal integrity engineers occasionally refer to this
as diode termination, but actually it is more properly called diode clamping because
the diodes clamp the reflections; they do not prevent them.
Diode D1 conducts when the high-going reflection exceeds Vdd by a diode
drop, and diode D2 conducts when the signal is more negative than ground by a
diode drop. Together these diodes bound the signal to (Vdd + Vd) and (Vss − Vd),
Vdd
Zo
V ne
Vfe
TX
RX
Rterm = Zo
Vtt
Vdd
Rterm = Zo
V ne
TX
Zo
Vfe
RX
Vtt
Figure 12.4 A proper parallel termination is formed with a single resistor equal to Zo that is connected to ground or Vdd.
144
Termination Strategies
Vdd
D1
Vne
Zo
Vfe
TX
RX
D2
Vtt
Figure 12.5 Clipping reflections with Schottky or low-capacitance, small-signal diodes.
where Vd is the drop across the diode when it is conducting the clamping current.
Because practical diodes include parasitic series resistance that changes in value
with the current passing through the diode, Vd is higher than the voltage needed to
turn the diode on. For instance, at room temperature a small signal switching diode
has a turn on voltage of about 0.6V, but the parasitic series resistance can cause the
clamped voltage to be closer to a volt.
The diode clamping action means that in systems where the receiver operates
from a power supply lower in voltage than the driver, D1 must be connected to the
driver power supply and not to the receiver power supply. Otherwise, diode D1
will conduct whenever the driver transmits a logic high, even when reflections are
not present. This static power dissipation can increase the junction temperature of
the driver and wastes system power.
Because connecting D1 to the driver supply cause clamping to occur at a voltage higher than the receiver supply, in dual supply systems the clamping voltage
may be not be low enough to prevent signal reflections from exceeding the maximum allowed for the receiver inputs. This is not a factor when the receiver supply
is a higher voltage than the driver supply, but in that case clamping at the higher
voltage will not be as effective. For instance, if the driver supply is 2.5V and the
receiver is 3.3V, D1 will not clip reflections unless they reach roughly 4V. The 3.3V
receiver is probably capable of withstanding this voltage, but the reflection sent
back down the line may not be as easily tolerated by the driver and will encourage
crosstalk to neighboring signal lines. For these reasons diode clamping is often not
as effective as actual termination in situations where the driver and receiver operate
at different voltages.
12.4.1 Diode Types
Diodes for signal clamping are either small signal switching diodes or, because of
their lower turn-on voltage and faster switching response, Schottky barrier diodes
[2, 3].
The diodes parasitic capacitance adds to the total receiver load capacitance.
Diodes with low junction capacitance (under 2 pF) and small charge storage times
(the diode transit time) are readily available, especially in surface mount packages,
and are preferred to minimize this undesirable parasitic loading.
12.5
AC Termination
145
A load line analysis can be performed to determine how the diodes on resistance affects the clamped voltage, but the effects are seen best with a circuit simulator. The simulations should include diode models from different manufacturers
because even for the same part number the circuit models often have different
characteristics. Therefore, simulations must show that the clamping is acceptable
with models from all the vendors supplying diodes for the build.
Although individual diodes can be used, dual diodes in three lead surface
mount packages specifically made for signal clipping are available that include
D1 and D2 with a common connection between them. The common lead connects to the signal trace and the other two leads connect to Vdd and ground. Some
examples of small signal silicon dual diodes in surface mount packages include the
MMBD1703, BAV99, MMBD4148SE, and the BAS40-04 (Schottky).
As part of the diode selection process, the circuit simulations should include
the measurement of the current through diode. The signal integrity engineer must
ensure that the clamping current is well within the diode current rating under all
switching conditions and environmental extremes.
12.5
AC Termination
The scheme shown in Figure 12.6 places a capacitor in series with the termination
resistor. This blocks the DC path so the network only dissipates AC power.
The acceptable value of blocking capacitor Cterm falls within a wide range. The
general requirement is for the capacitors reactance to be very much smaller than
the resistance of Rterm. In this way the capacitor acts as an AC short, connecting
Rterm directly to Vterm (such as to ground, as is shown in the figure). However, if it
is made too large, excessive time is required to establish the proper bias across the
capacitor, adversely affecting the received signal amplitude.
This termination is best used with clocks or other repetitive signals, and is usually not suited for data buses or address lines where the frequency content of the
data stream varies. Very low frequencies corresponding to the back-to-back transmission of many same sense bits require a large capacitor, but such a large value is
not appropriate when the signal frequency is high (corresponding to an alternating
1/0 pattern, for instance) because of the prolonged time required for the capacitor
Vdd
V ne
Zo
Vfe
TX
RX
Rterm = Zo
Cterm
Vtt
Figure 12.6 With AC termination TX only dissipates AC power. Cterm’s reactance is much less than
Rterm.
146
Termination Strategies
to charge to the bias level. For this reason using AC termination on data buses results in significant data-dependent jitter [4] and is usually avoided.
Simple equations exist to calculate Cterm from the signal rise time [5, 6], but
these are misleading because they underestimate the capacitance needed for fast
rise time, low-frequency signals. Equation (12.4) avoids this by providing an estimate based on the signal frequency.
Cterm =
3
f × Zo
(12.4)
The equation computes the capacitance with a reactance 1/20th of Zo at the
frequency f. The capacitor reaches its steady state bias level within 15 back-to-back
transitions. By that time the circuit is properly terminating the signal, but pulses
arriving before this time may not be.
For instance, Cterm is 60 nF when Zo is 50Ω and f is a 1-MHz clock. The capacitor has reached the steady state bias point by the 15th clock pulse (15 μs).
The calculated value should be used as a starting point when performing circuit
simulations of the termination. The simulations ought to include the interconnect
parasitics (such as the via inductance) that are in series with the RC network since
these increase with frequency, diminishing the ability of Cterm to act as an effective
connection for Rterm.
Discrete resistors and capacitors are usually used to create the AC terminator,
but integrated RC networks are also available. As demonstrated in the Problems,
the capacitance of these networks is small, which makes them most useful in terminating clocks with frequencies in the hundreds of megahertz range.
12.6
Topologies
In Figure 12.7 the wiring between surface mount integrated circuits U1 and U4 is a
single load point to point connection, while the connection from U2 to U3, U5, and
U6 is an example of point-to-point multiload topology. The connection between
integrated circuits U7 through U10 is a multidrop topology.
These topologies have different electrical characteristics and, as we will see,
different response to various types of termination.
12.6.1 Single-Load Point-to-Point Termination
The series and parallel termination of a point-to-point single-load connection is
shown in Figure 12.8.
The series termination illustrated in Figure 12.8(a) shows a long transmission
line T2 connecting a capacitive load CL to a series terminating resistor Rs. Ideally, Rs is connected directly to the driver, but in general practice a length of trace
(shown as transmission line T1) is used to make the connection. This pin escape
trace is necessary because the physical size of Rs usually makes placing it immediately next to the driver pin impossible. This is especially true with large, fine-pitch
multipin ASICs where the pins (or balls if in a BGA package) are arrayed in several
12.6
Topologies
147
U1
U2
U3
U4
U5
U6
U9
U10
U7
U8
Figure 12.7 Circuit board artwork showing point-to-point topology from U1 to U4; single load
point-to-point from U2 to U3 and U5 – U6; multidrop from U7 to U8 through U10.
rows. In that case T1 is long enough to escape the pin field and to place the signal
in an open area where Rs can be mounted.
In lower-density boards the pin escape trace T1 is often routed as microstrip,
but high-performance, high-density boards will frequently escape with stripline. In
either case, the impedance of T1 (Zo1) is often not the same as that of T2 (Zo2). If
T1 is too long, its impedance will determine the current launched by the driver and
reflections will occur even if Rs properly matches T2. For this reason the length of
T1 should always be short compared to the signal rise time, and it should always
be less than the length of T2.
148
Termination Strategies
Zo1
TX
Zo2
RS
T1
T2
CL
(a)
TX
Zo2
Zoo3
T2
T3
CL
Rterm
Vterm
(b)
Figure 12.8 (a) Series and (b) parallel termination of a single point-to-point load. Capacitor CL
represents the receiver input capacitance.
The parallel termination illustrated in Figure 12.8(b) shows how the long connecting trace T2 directly connects to CL and how the termination connects through
transmission line T3. Although a Thevenin termination is pictured, it actually represents any of the parallel termination schemes we have already discussed.
In general, the termination is usually some distance from the load and trace T3
may be longer than T2. This fly-by termination is especially likely if the board is
very densely populated because often space is not available to place the termination
immediately at the pin of the receiving device. In this case T3 allows the signal to
fly by the load and be terminated by the resistor, which is placed in an open area.
The length of T3 is not critical provided T2 and T3 have the same impedance.
By making Rterm equal in value to Zo3, transmission line T3 is properly terminated
and for any length it appears to CL as a resistor of value Rterm. This means T2 and
T3 must be formed on the same routing layer.
Although T3 need not be short, very long lengths must be analyzed carefully,
especially when signaling at high data rates. Very long lengths of trace will be lossy
at high frequencies, which invalidates the idea that any length of T3 is the same as
connecting Rterm directly to the load. Also, crosstalk is more likely to occur on long
lengths since there are more opportunities for T3 to be exposed to coupling from
other signals. Regardless of T3’s length, it should be considered a critical net and
carefully scrutinized for crosstalk.
12.6.2 Multiple Load Point-to-Point Termination
Multiload point-to-point termination schemes are shown in Figure 12.9.
12.6
Topologies
149
R S1
T1
R S2
TX
T2
R S3
T3
C L3
C L2
C L1
(a)
Rterm1
T1
Rterm2
TX
T2
Rterm3
T3
C L3
C L2
C L1
Vterm
(b)
Figure 12.9 (a) Series. (b) Parallel multiload point-to-point termination schemes.
Series termination is shown in Figure 12.9(a), and parallel termination is shown
in Figure 12.9(b), where we see that three loads are connected by three transmission lines to the driver, and in both cases the driver must simultaneously launch
current down the three lines. This effectively places them in parallel. For instance,
if the impedance of each line is 60Ω the driver experiences a load of 20Ω. Unless
the driver has a very low output impedance, it may not be capable of launching a
voltage high enough to reach the receiver switch point into such a low impedance.
With the series termination scheme, reflections from the impedance mismatch
at the load will increase the received voltage, provided that the load impedance is
very high. This means that the driver need not launch as high a voltage and can be
weaker than in the parallel terminated case.
However, with this scheme the relative lengths of T1, T2, and T3 can become
important. If the transmission lines are the same length, the reflections from all of
the lines simultaneously arrive at the near end, and the waveform is the same at
each load. However, if the lengths differ, the reflections arrive back at the driver at
different times, which then combine in complex ways before they are relaunched to
the loads. This alters the wave shape at each load. An example of this for the series
termination scheme is shown in Figure 12.10.
In this example the driver has a nominal impedance of 12Ω and is connected
to a 3.3-V source. The driver “fans out” a common clock to three loads that are
located at different distances. Series resistors RS1 through RS3 are each 22Ω.
150
Termination Strategies
3.5V
3.0V
2.5V
2.0V
1.5V
1.0V
0.5V
0.0V
10 ns
15 ns
(a)
20 ns
25 ns
10 ns
15 ns
20 ns
25 ns
(b)
Figure 12.10 Series-terminated clock waveform at the end of the 3-inch (7.6-cm) trace (a) when
all traces are the same length and (b) when the trace lengths differ (plotted with the same vertical
scale).
Figure 12.10(a) shows the waveform at load capacitor CL3 when traces T1, T2,
and T3 are all 3 inches (7.6 cm) in length. The rising and falling edges transition
smoothly, without kinks (the signal is monotonic), and then high- and low-going
overshoot is well controlled.
In contrast, Figure 12.10(b) shows the voltage at CL3 when transmission line
T1 is 1 inch (2.5 cm), T2 is 2 inches (5.1 cm), and T3 is 3 inches (7.6 cm) in length.
The effect on the load of the unequal arrival time of the pulses at the near end
is evident. This waveform is not acceptable for a clock. The falling edge glitches
up to over 0.75V, and the negative-going overshoot is pronounced. To correct this
waveform, the traces should all be length matched, as shown in Figure 12.10(a), or
a clock fan-out device, a specialty integrated circuit made to distribute low skew
copies of an input signal, could be used.
12.7
Multidrop Lines
As shown schematically in Figure 12.11, rather than using point to point connections, a multidrop connection can be used to send a signal from a driver to multiple
loads.
The three load capacitors CL1 through CL3 are connected to the driver by taping into a transmission line. Each of these lines may be very long compared to the
signal rise time, or (more commonly) their electrical length may be comparable to
the rise time. Without proper termination and attention to length matching, signals
on multidrop lines are particularly likely to be nonmonotonic and to experience
multiple reflections.
12.7
Multidrop Lines
151
RS
TX
T1
T2
C L1
TX
T1
T3
C L2
T2
C L1
C L3
T3
C L2
C L3
Rterm
Vterm
Figure 12.11 Multidrop connection. Either source series termination (using Rs) or one of the parallel schemes (Thevenin Rterm and Vterm as shown here) may be used.
12.7.1 Response When Signal Rise Time Is Very Short
The response of the circuit when the signal rise time is much smaller than the electrical length of the line is illustrated in Figure 12.12. Each transmission line is 50Ω
and 6 inches (15.2 cm) in length. Load capacitors CL1 through CL3 are each 2 pF
and represent the capacitance of a receiver.
4.5V
4.0V
3.5V
3.0V
2.5V
C L1
2.0V
1.5V
C L1
1.0V
C L2
C L2
C L3
0.5V
C L3
0.0V
0s
5 ns
(a)
10 ns
0s
5 ns
(b)
10 ns
Figure 12.12 Response to the circuit shown in Figure 12.11 with (a) series termination and (b)
parallel termination. Load CL1 is the last to fully switch when series terminated, but it is the first when
parallel terminated.
152
Termination Strategies
It is evident from Figure 12.12(a) that when the line is series terminated, a
plateau voltage is created at each load as the incident wave passes by. In this case,
the plateau is not quite equal to half the value of Vdd, the power supply feeding the
driver.
In general, if the plateau voltage is greater than the receiver switching threshold, the receivers will detect the voltage as a valid logic level and will switch on the
incident wave. However, incident wave switching is not occurring in this particular
case because the plateau voltage does not exceed the receiver switch point.
Instead, each of the loads must wait until the reflection created at the open circuit at the very end of the line arrives and increases the plateau to a higher voltage.
Since the reflection is created at the furthest load, CL3 is the first to switch. We can
see in Figure 12.12 how each of the remaining loads switch in reverse order as the
reflection travels from CL3 toward CL1.
In contrast, the response when parallel terminated [appearing in Figure
12.12(b)] shows that there is no plateau and reflections are not required for the
signal to reach the switch point. In this case, the load closest to the driver (CL1)
switches before CL3.
Although the parallel termination dissipates more power (including power in
the driver), it gives the fastest overall response because the signal is high enough
to exceed the receiver’s switch point as it first passes by rather than having to wait
for the reflection.
This timing difference is evident in Figure 12.12. For instance, if we use 2.4V
as the receiver switch point, we can see that in this example load CL1 reaches that
voltage at a time of about 1.2 ns when parallel-terminated and at roughly 5.7 ns
when series-terminated. However, the load CL3 reaches 3.4 ns for both the series
and parallel termination schemes.
12.7.2 Response When Signal Rise Time Is Comparable to the Transmission
Line Delays
In Chapter 6 we saw that if the line was longer than half the signal rise time, the line
components acted as distributed elements. We can look at this in the reverse and
say that if the signal rise time is longer than twice the electrical length of the line
than the elements appear as a single lump. For instance, a transmission line having
a 0.5-ns electrical length would appear as a single RLC lump if the signal rise time
was 1 ns, but it would appear more as a distributed circuit if the rise time was less
(such as 750 ps).
We will use this insight to analyze the effects of loads evenly distributed along
a transmission line when the rise time is comparable to the electrical length of the
line.
A driver source series terminated to a 12 inch (30.5 cm) 65Ω transmission line
is shown in Figure 12.13. Without loads the total one way delay of this line is 2.1
ns. A receiver (represented by load capacitors CL1 and CL2) are placed midway and
at the end of the line. A 65Ω resistor Rs is used to impedance match the driver to
the transmission line.
A 3.3-V driver with an on resistance much lower than 65Ω launches a fast rise
time signal down the transmission line. The dotted curve in Figure 12.14 shows
that the voltage Vne has an initial plateau of 1.65V. Since this is one half of the
12.7
Multidrop Lines
153
6”
(1.05 ns)
6”
(1.05 ns)
T1
T2
Zo =65Ω
Zo =65Ω
RS
3.3V
TX
65Ω
Vne
C L1
3.5 pF
C L2
3.5pF
Figure 12.13 Capacitive loads CL1 and CL2 placed along and at the end of a series terminated transmission line. Signal rise time determines T1’s apparent impedance.
3.5V
3.0V
2.5V
1.65V
2.0V
1.5V
1.57V
1.0V
0.5V
0.0V
0s
2 ns
4 ns
6 ns
8 ns
10 ns
Figure 12.14 Voltage measured at Vne when the signal has a very sharp rise time (dotted curve) and
when it is longer (solid curve). The two voltages show that the impedance is not the same.
driver voltage this confirms the transmission line impedance is 65Ω. The negative
going pulse caused by CL is clearly visible.
The solid curve shows the response when the signal rise time is extended to 2.5
ns. The plateau is still present, but it has been shifted down to 1.57V. As we first
saw in Chapter 6, the voltage divider principle can be used to determine that the
transmission line impedance has fallen from 65Ω to 59Ω. The negative-going pulse
caused by load capacitor CL1 that is so evident with the sharp-edged pulse is no
longer visible.
Lengthening the signal rise time causes these changes because now the line’s
distributed capacitors and inductors become indistinct and blend together, appearing as a single inductor and capacitor. Capacitor CL adds to the transmission line
capacitance, increasing it, but it does not change the transmission inductance.
The transmission line inductance (L) and capacitance (C) affects the transmission line impedance (Zo) as is shown in (12.5) (which was first presented in Chapter 6).
Zo =
L
C
(12.5)
154
Termination Strategies
Since the capacitance has increased and the inductance has not changed, (12.5)
shows how adding CL1 lowers the transmission line impedance. In fact, as is evident by the absence of the negative-going pulse at the end of the plateau region, CL1
no longer shows up as a separate element.
Usually the loads in CMOS system are mostly capacitive, making it possible to
ignore the inductance. Doing this lets us calculate the loaded impedance, ZL (12.6)
[6, 7].
ZL =
Zo
1+
CL
Co
(12.6)
where Ls and Cs are the self-inductance and capacitance of the transmission line per
unit length, and CL is the load capacitance evenly distributed along the transmission
line length.
Similarly, distributed loads increase the effective delay of a transmission line.
The loaded delay time (tpdL) is given in (12.7).
tpdL = tpd 1 +
CL
Co
(12.7)
Often for a circuit board trace we know Zo and tpd and not Co. Simultaneously
solving the equations for delay and impedance presented in Chapter 6 lets us find
Co in terms of Zo and tpd, presented here:
Co =
tpd 2
tpd 2 × Zo2
(12.8)
Using Figure 12.13 as an example, assume the 12-inch-long (30.5-cm), 65Ω
line has an end-to-end delay of 2.1 ns without loads. Each 6-inch segment therefore
has an unloaded delay of 1.05 ns. From (12.8), Co for each segment is 16.2 pF. Using this and the value of CL (3.5 pF) in (12.6), we find that ZL is 59Ω, matching the
results in Figure 12.14. By using (12.7), we find that the delay for each segment has
increased from 1.05 ns/inch to 1.2 ns/inch (430 ps/cm to 480 ps/cm).
As the Problems illustrate, because the self-capacitance of low-impedance traces is so high, their impedance and delay are less affected by small value capacitive
loads.
12.8 Stubs and Branches
In Figure 12.11 the loads are shown as being connected directly to the transmission
line, but usually a short length of transmission line (a stub) is used to make the connection. This is shown schematically in Figure 12.15.
By using the definition developed in Chapter 6, we can say that the stub would
be considered long if its electrical length is greater than half the signal rise time.
12.8
Stubs and Branches
155
RS
TX
N1
T1
N2
T2
T3
T4
C L1
T5
C L2
C L3
Figure 12.15
– T3).
Stubs T4 and T5 connect the load capacitances to the main transmission line (T1
For instance, if the signal rise time is 1 ns, transmission lines T4 and T5 would
appear to be long if their electrical lengths were 0.5 ns or greater.
If the electrical length of T4 and T5 is small, then the capacitance of the stub
is found with (12.8) and is added to the load capacitance as we did in the previous
section. This increases the load capacitance and creates a larger impedance discontinuity at the point where the load connects to the line. However, because the stub
is short, incident and reflected waves will not appear along its length.
12.8.1 Branches
When the trace connecting the load to the main transmission line is electrically long,
it acts as a branch rather than a stub. The branch is then in parallel with the main
transmission line, which lowers the impedance where the branch connects. Figure
12.15 is redrawn as Figure 12.16 to emphasize this parallel connection. The net can
be series- or parallel-terminated.
It is evident from Figure 12.16 how an incident wave exiting T1 would simultaneously send energy down the parallel combination of T2 and T4. In this particular
case T2 and T4 are each 50Ω, so the impedance experienced by the incident wave
when it reaches N1 is 25Ω. Unless terminated at the load, the wave launched down
T4 will reflect when it reaches load capacitor CL1. That reflection will be rereflected when it reaches node N1, because from the point of view of this wave, the
50Ω transmission lines T1 and T2 are in parallel, and node N1 has an impedance
of 25Ω. This same process occurs at node N2 with transmission lines T3 and T5.
Depending on the relative lengths of T2 through T4, the reflected energy can create
complex-looking waveforms at the loads, and the waveforms at each of the loads
can be quite different.
We can see from this how ineffectual series termination will be at controlling
these reflections. In fact, the best signal integrity will occur when each of the loads
is individually terminated, but even so, the impedance mismatch between the 50Ω
transmission line T1 and the apparent 25Ω impedance appearing at node N1 (and
node N2) will still give rise to reflected energy.
This can be solved by lowering the impedance of T1 to 25Ω, matching T1 to
the impedance at N1, and eliminating any reflections from the incident wave when
156
Termination Strategies
Rterm3
T3
T2
C L3
N2
Rterm2
RS
TX
T5
T1
N1
C L2
Rterm1
T4
C L1
Vterm
Figure 12.16 Electrically long transmission lines T4 and T5 are in parallel with transmission lines T2
and T3. This net may either be series-terminated with Rs or parallel-terminated using Rterm.
it reaches N1. Provided that CL1 is properly terminated, reflected energy will not be
present to travel through T4 back to N1, so there is no difficulty with the parallel
impedance of T1 and T2 being so low (under 17Ω in this case).
These results are shown in Figure 12.17 for 50 transmission lines. Figure
12.17(a) shows the voltage across load capacitor CL1 and CL3 when a 15Ω driver
is series-terminated with 50Ω (Rs = 50Ω). In Figure 12.17 parallel termination is
not used. The poor signal quality is evident.
Figure 12.17(b) shows the load voltages with Rs removed and 50Ω, 1.65-V
Thevenin terminators are connected to CL1, CL2, and CL3. The signal quality is
greatly improved, but the effects from the impedance mismatch at N1 and N2 are
still apparent.
As is shown in Figure 12.17(c), further improvement can be obtained by lowering the impedance of T1 to 25Ω and retaining the 50Ω, 1.65V Thevenin terminators at each load. Lowering the impedance eliminates the mismatch at node N1,
and the Thevenin termination reduces reflections from the loads.
Generally, this technique only works for unidirectional signals because it is
usually not possible to properly impedance match all branches in all directions. For
instance, replacing CL1 with a second transmitter would require that the impedance
of T4 be lowered to the parallel combination of T1 and T2, but we have already
seen that for proper operation when TX drives, transmission line T1’s impedance
has to be lowered to the parallel combination of T2 and T4.
Although Thevenin termination was used in Figure 12.17, any of the parallel
termination techniques previously described can be used to terminate the branches.
The fundamental point is that parallel termination prevents (or at least minimizes)
reflections from being created by the loads, so the impedance mismatch created by
12.9
Main Points
157
3V
2V
1V
0V
(a)
3V
2V
1V
0V
(b)
3V
2V
1V
0V
0s
5 ns
10 ns
15 ns
(c)
20 ns
25 ns
Figure 12.17 (a) Load capacitor CL1 voltage when Rs = 25Ω, Rterm not present and T1 – T5 = 50Ω.
(b) Rs removed, Rterm = 50Ω, Vterm = 1.65V. (c) Rterm = 50Ω, Vterm = 1.65V, T1 = 25Ω.
the parallel combination of transmission lines at a branch is not as significant as it
would be if the loads were not terminated. This means that because diode clamps
clip the reflections and do not prevent them from being created, they usually are
not a useful substitute in this situation for one of the other parallel terminations.
12.9
Main Points
•
Transmission lines may be terminated at the source (source series termination) or at the load (parallel termination).
•
Source series termination raises the impedance of the driver and can prevent
reflected energy from rereflecting from the generator.
•
Parallel termination impedance matches a high-impedance load to a lowerimpedance transmission line. There are several variants, and circuit simulation should be used to determine which is best suited in a given application.
•
Capacitive loads connected to a transmission line will either appear as a
sharp discontinuity (if the signal rise time is very fast) or will combine with
the transmission line capacitance to lower it.
•
Connections made to a transmission line are either stubs (when short) or
branches (when they are long). Their effects are very different.
158
Termination Strategies
Problems
Answers to these problems are available on the Artech House Web site at http://
www.artechhouse.com/static/reslib/thierauf/thierauf1.html.
12.1 A parallel termination uses 120Ω, ± 5% resistors for R1 and R2 (shown in
Figure 12.3). What is the likely impedance of the trace that this terminates?
12.2 Using the same setup described in Problem 12.1, determine if ± 5% tolerance
resistors are adequate or if resistors with tighter tolerance are necessary.
12.3 Compare the performance of a 60Ω Thevenin termination shown in Figure
12.3 that uses a ± 5% resistor with the parallel termination illustrated in
the previous problem.
12.4 Find the proper values for R1 and R2 (shown in Figure 12.3) to terminate
a 75Ω coax cable to 1V when Vdd is 5V.
12.5 A system uses logic devices that have an input threshold of 2.4V. To guarantee that this exceeds the receiver threshold, the value of Vterm in Figure
12.3 is set to be 3V. Rterm must be no less than 130Ω to avoid exceeding the
current handling ability of the driver. Find the value of R1 and R2 when
Vdd is 5V.
12.6 Determine the difference in power dissipation in Figure 12.3 when the bus is
held low if R1 and R2 are both 100Ω. Assume that Vdd is 3.3V and Rds_on
is 12Ω.
12.7 What is the minimum value capacitor that should be used to AC terminate
a 65Ω trace that is carrying a 33-MHz clock?
12.8 An integrated RC network used for AC termination has a resistance Rterm
of 50Ω and a capacitance Cterm of 470 pF. What is the lowest frequency that
this network can reasonably be expected to terminate?
12.9 What is the proper value for RS1, RS2, and RS3 in Figure 12.9 when T1 has
an impedance of 50Ω and the impedance of T2 and T3 is 65Ω? Assume
that the signal rise time is 50 pS, the load capacitors are all 2.5 pF, and the
transmission lines all have a delay of 2 ns.
12.10 Using the same information as in the previous problem, what should be the
values of Vterm and Rterm1 through Rterm3 when Vdd = 3.3V and the receivers
V
represented by CL1 through CL3 switch at dd ?
2
12.11 Find the loaded impedance of the transmission line shown in Figure 12.13
when Zo is lowered to 40Ω. How does this change compare to the 65Ω
example given in Section 12.7.2?
12.12 What values should be chosen for transmission line T1 and resistor RS and
Rterm1 through Rterm3 in Figure 12.16 when T2, T3, and T5 are 50Ω and
T4 is 65Ω?
12.13 Three state drivers are used to drive data between nodes with the topology
shown in Figure 12.16. Assume that reducing power is important in this application, and to facilitate this, the bus will be made inactive (set to the highimpedance state) at frequent intervals. If the transmission lines are 50Ω and
Vdd is 3.3V, what form of parallel termination (shown in Figure 12.3) is the
best choice to reduce power?
Problems
159
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
Seams, J., “A Comparison of Resistive Terminators for High Speed Digital Data Transmission,” High Frequency Electronics, October 2005, pp. 18–26.
Hewlett Packard Corp., “Applications Bulletin 14: Waveform Clipping with Schottky Diodes,” undated.
Hewlett Packard Corp., “Applications Bulletin 15: Waveform Clamping with Schottky
Diodes,” March 1989.
Ethirajan, K., and J. Nemec, “Termination Techniques for High-Speed Buses,” EDN, February 16, 1988, pp. 135–145.
Vu, M., “Choose Termination and Topology to Maximize Signal Integrity and Timing,”
EDN, October 24, 1996, pp. 95–98.
Bakoglu, H. B., Circuits, Interconnections, and Packaging for VLSI, Reading: AddisonWesley, 1990.
Fairchild, Inc., ECL Data Book, 1977.
C H A P T E R 13
Differential Signaling
13.1
Introduction
The previous chapters discussed single-ended signaling, where a signal voltage is
sent down a single transmission line and is compared with a reference voltage to
determine its logical value. This scheme is shown in Figure 13.1(a). It is apparent
how a voltage difference between the ground of the driver and the ground of the
receiver (Vnoise) can reduce the received noise margin, leading to false triggering.
HSTL [1] and SSTL [2] are popular examples of single-ended signaling.
This contrasts with differential signaling, where a differential driver simultaneously sends a signal and its complement down two transmission lines (a differential transmission line or diff-pair). As shown in Figure 13.1(b), the differential
driver switches one output from a logic low to high while simultaneously switching
the complementary output from high to low. This differential signal is detected
by a differential amplifier (a diff-amp), which uses the voltage difference between
the two signals (and not the voltage measured to ground) to determine if it is a
logic high or low. Common examples of differential signaling include LVDS [3],
Infiniband [4], and RapidIo [5].
One of the diff-pairs is usually called the negative terminal even when the signaling is done with positive voltages. For instance, LVDS signals typically switch
between +1.0V and +1.4V (and is centered around a common mode voltage of
1.2V [6]), but the driver outputs (DO+ and DO) and receiver inputs
(RI+ and RI−) are referred as the “positive” and “negative” terminals.
In comparison to single-ended signaling, differential signaling:
•
Requires more signal traces and I/O pins;
•
Retains good noise immunity even when signaling with lower voltages;
•
Requires different termination techniques;
•
Is more immune to differences in voltages between the driver and receiver
grounds;
•
Is more immune to noise coupled on the signal traces (provided the noise
couples equally to them both).
161
162
Differential Signaling
Vtt
RI
RI
+
−
+
Vtt
Vnoise
(a)
RI−
Vcm
DO+
RI+
DO−
RI−
RI+
+
−
(b)
Figure 13.1 (a) In single-ended signaling the receiver switch point (Vtt) is with respect to the local
ground. (b) With differential signaling the crossing of complementary signals determines the switch
point.
13.2 What Are the Electrical Characteristics of Differential Signaling?
A diff-amp receiver amplifies the difference in voltage appearing on its inputs (RI+
and RI−) and rejects voltages that simultaneously appear on both inputs. This is
why noise that couples equally to both inputs is ignored by the receiver. In fact, to
get the most benefit from differential signaling, we will intentionally exploit this
characteristic by deliberately routing the two traces forming the diff-pair so that
noise is encouraged to couple equally to both traces.
Taking the receiver ground connection as a reference, the difference in the
ground voltages between the transmitter and receiver appears to the receiver as a
noise voltage that adds or subtracts from the incoming signal. However, because
they do not use ground as a signal reference, the switch point of differential receivers is not affected by this type of common mode noise.
However, there are limits in the common mode voltage over which the circuit
will properly operate. The receivers’ common mode range specifies the maximum
voltage that the signal may reach (measured with respect to the ground local to the
receiver) before the transistors within the diff-amp become unbiased and it loses
the ability to differentiate valid differential signals. This is a critical parameter to
assess when performing signal integrity simulations and when making laboratory
measurements.
13.2 What Are the Electrical Characteristics of Differential Signaling?
163
13.2.1 How Is a Differential Pair Observed?
A probe with a ground clip is used to make single-ended oscilloscope measurements
in the laboratory. The ground clip sets the reference for the measurement by specifying the location where zero volts is defined. This is also how SPICE simulations are
customarily observed, since, unless told otherwise, SPICE calculates and displays its
results with respect to node 0, the system ground.
In contrast, a differential probe is used to measure the differential signal. The
displayed signal is truly differential: The difference between RI+ and RI− appearing
in Figure 13.1 is measured rather than its value relative to ground. An example of
a differential signal is shown in Figure 13.2.
We see that each of the signals swing between 0V and 1V with respect to
ground, but the difference signal swings from +1V to −1V.
Because an oscilloscope only allows a few transitions at a time to be displayed,
it is difficult to observe the interaction between data bits occurring at significantly
different places in a long stream of data. Setting the oscilloscope time base to simultaneously display all the bits does not provide the resolution necessary to distinguish a single pulse. Rather than an ordinary oscilloscope, a device that captures
and then overlays successive pulses (bits) is used. One example of a device that can
do this is the digital signal analyzer such as the one shown in Figure 13.3.
13.2.2 What Is an Eye Diagram?
A digital signal analyzer creates an eye diagram [8–12] such as that shown in Figure
13.4 by capturing a data stream and overlying successive pulses. Ideally, all of the
rising edges should exactly overlay, as should all of the falling edges. The space in
between should be totally clear (void of any transitions). The data mask defines the
minimum acceptable width and height of this opening and is a specification that the
link must meet for proper operation.
1V
0V
−1V
(a)
1.0V
DO−
0.5V
DO+
0.0V
(b)
Figure 13.2 (a) The differential signal (DO+) – (DO−) swings between ± 1V even though (b) when
measured respect to ground the signals switch from 0 to 1V.
164
Differential Signaling
Figure 13.3 A digital signal analyzer such as this Tektronix DSA 8200 [7] is used to display a differential data stream. (Courtesy of Tektronix, Inc. Used with permission.)
Latest falling edge
All rising edges lay within
this region
Highest Vih
Earliest falling edge
Timing
uncertainty
(jitter)
Lowest Vih
Mask
Earliest rising edge
Latest rising edge
Highest Vil
All falling edges lay
within this region
Lowest Vil
Figure 13.4 The earliest and latest appearances of the rising and falling edges and their amplitudes are
simultaneously displayed in this eye diagram. (From: [13]. Used with permission.)
In reality jitter will prevent the pulses edges from perfectly lining up, and reflections and losses will prevent each pulse from having the same logic low (Vil) and
high levels (Vih). These things reduce the horizontal and vertical eye opening and,
if severe enough, can prevent proper operation.
13.2.3 What Is Intersymbol Interference?
We saw in Figure 8.2 how transmission line losses cause a pulse to shrink in height
(amplitude) and for its base to grow in width. This widening at the base is particularly serious because if it grows large enough, successive pulses (symbols) will blend
together, making it difficult to distinguish between each one. Symbols can also
13.3 What Are the Electrical Characteristics of a Differential Transmission Line?
165
become distorted when they combine with residual energy remaining on the line
caused by reflections from earlier symbols.
Any such interference between symbols [intersymbol interference (ISI)] alters
the shape of the waveform appearing at the receiver [14–16] and is easily observed
with an eye diagram. Fortunately, signal processing techniques have been developed that allow proper signal reception of very high-speed signals, even on lines
where ISI and losses are high.
13.2.4 How Are the Effects of Loss Corrected?
We have seen in previous chapters how transmission line losses increase as frequency increases, and we also saw that a signal with fast rise times have a high-frequency content. Since high-data rate signaling requires sharp, fast rise time pulses,
it is apparent how transmission line losses can become significant when signaling at
high data rates. In fact, even in the absence of crosstalk or other kinds of noise, in
high-speed signaling it is quite likely for frequency dependent losses to distort the
pulses enough to prevent the link (the hardware and software connection between
a transmitter and receiver) from operating properly.
It is very common to use an equalizer on high-speed links to improve the quality of the received signal [11, 12, 14, 15]. The equalizer can either be a passive
circuit [13, 17] or a precompensation circuit built into the transmitter. With the
precompensation technique [10, 12, 18–20], the transmitter monitors the data pattern that it is sending and intentionally predistorts the signal to compensate for the
transmission line loss.
When done properly, the predistortion applied by the transmitter and the distortion caused by the transmission line interact in such a way as to produce in a
much cleaner signal at the receiver. An example of the benefits of a properly equalized net is shown with the eye diagram appearing in Figure 13.5.
The eye in Figure 13.5(a) is difficult to find and is an example of unacceptable signal integrity. Activating equalization at the transmitter as shown in Figure
13.5(b) dramatically reduces jitter and improves ISI. The eye is clearly defined and
is large enough to permit reliable link operation.
13.3 What Are the Electrical Characteristics of a Differential
Transmission Line?
In Chapter 9 we saw how two traces that were carrying signals exactly out of phase
with one another operated in the odd mode. Because this describes the way in which
ideal differential signals switch, differential signaling operates in the odd mode.
This suggests that to form a diff-pair the odd-mode impedance should be set to
the desired impedance. However, as shown in (13.1), the differential impedance is
actually twice the odd-mode impedance. This profoundly affects the way in which
differential transmission lines are created and terminated.
Zdiff = 2Zoo
(13.1)
166
Differential Signaling
(a)
(b)
Figure 13.5 Measured eye diagram showing the link (a) with no equalization and (b) with equalization. (From: [13]. Used with permission.)
13.3.1 Why Is the Differential Impedance Twice the Odd-Mode Impedance?
Figure 13.6 shows a 50Ω differential driver connected to a pair of traces that have
an odd-mode impedance of 50Ω. From (13.1) this setup has a differential impedance of 100Ω, and we will prove this by calculating the voltages and currents
launched by the driver.
During the time t, the driver is transmitting a logic high value by sending a high
voltage on the DO+ line and a low voltage on the DO− line. To find the differential
impedance, we must determine the values of these voltages and of the launched current. With this we will use Ohm’s law to find the impedance.
The termination resistance Rterm placed across the diff-pair converts the
launched differential current into a differential voltage that is detected by the diffamp receiver. The received voltage swings about a common-mode voltage of 0.5V.
The diff-amp input terminals have very high impedance and they draw no current.
By using the voltage divider principle as described in Chapter 6, we might conclude that the 1-V, 50Ω driver would launch a 0.5-V signal down the 50Ω transmission line. In fact, 0.75V is launched. To understand why this is so, we redraw the
circuit as shown in Figure 13.7, diagramming the common-mode voltage and the
current flow. The diff-amp receiver inputs draw no current and so can be ignored.
13.3 What Are the Electrical Characteristics of a Differential Transmission Line?
167
Differential
driver
50Ω
1V
Zoo = 50Ω
DO+
+
RI+
R term
RI−
+
Vout
−
DO−
50Ω
0V
Zoo = 50Ω
RI−
0.75V
RI+
0.25V
Vcm = 0.5V
Vout
t
Figure 13.6 A 50Ω differential driver connected to a differential receiver. The signal appearing
across Rterm swings 0.25V above and below a 0.5-V common-mode voltage. The receiver output
shown at the bottom of the figure swings rail to rail.
Differential
driver
50Ω
Zoo = 50Ω
RI+
DO+
DO−
VDO+
+
0.5V
+0.25V
RT
Vcm
+
+
−0.25V
RT
0.5V
0.5V
VDO−
50Ω
RI−
DO−
Zoo = 50Ω
Figure 13.7 Differential driver connected to a differential pair. The outputs switch about the common-mode voltage Vcm. Voltages on DO+ and DO− are measured with respect to Vcm.
The circuit consists of a 0.5-V supply voltage around which the outputs switch
(the common-mode voltage, Vcm), and two 0.5-V supplies and 50Ω resistances
connected to the DO+ and DO− outputs. The termination resistor Rterm shown in
Figure 13.6 has been separated into two equal valued resistors called RT. They are
connected together to the common-mode voltage supply.
When driving a logic high, the common-mode supply adds to the supply connected to the DO+ pin. When measured with respect to ground, DO+ is a 50Ω resistance in series with a 1-V source. Similarly, the supply connected to the DO− pin
168
Differential Signaling
subtracts from the common-mode voltage, making it a 50Ω resistance in series with
a 0V source. This validates the model; the DO+ and DO− pins have the correct
50Ω impedance, and with respect to ground, they switch between the correct voltage levels (0 and 1V).
With this circuit arraignment the voltage divider principle can be applied to
find the launched voltage. Since the driver and transmission line impedances are
both 50Ω, the voltage on the DO+ and DO− pins will each be half the switched
voltage. As shown, this is ±0.25V when measured with respect to the commonmode voltage.
Our analysis to this point is illustrated in Figure 13.8. For simplicity only the
DO+ signal is shown. As expected the launched voltage is 0.25V higher than the
common-mode voltage. Ohm’s law can now be used as was done in Chapter 6 to
find that 5 mA is sourced by DO+. The DO− analysis is the same, noting that the
pin is 0.25V lower than the common-mode voltage. That driver sinks 5 mA from
the common-mode supply.
The arrows in Figure 13.7 show that the currents from the DO+ and DO− outputs flow in the opposite directions. Because they are the same value, this makes
the net current in the common-mode connection zero, and the connection can
be removed without disturbing the circuit operation. This has been done in Figure 13.9, where we see the voltages of the two outputs measured with respect to
ground rather than to the common-mode voltage. The DO+ output is 0.25V higher
the 0.5-V common-mode voltage, making it 0.75V with respect to ground. The
DO− output is 0.25V lower than the common-mode voltage, making it 0.25V.
The voltage difference between the two outputs is the differential output voltage and is given in:
Vdiff = VDO+ − VDO − = 0.75 − 0.25 = 0.5V
(13.2)
All the background information is now available to find the differential impedance; the differential voltage is 0.5V and the current is 5 mA. Ohm’s law determines
the relationship between the transmission line impedance, voltage, and current. In
this case the differential impedance is 100Ω, as given in:
Zdiff =
Vdiff
Idiff
=
0.5V
= 100Ω
5 mA
(13.3)
Zoo = 50Ω
50Ω
DO+
RI+
VDO+
5mA
+
0.5V
+0.25V
RT
I = 0.25V/50Ω = 5 mA
Figure 13.8 A 0.25-V, 5-mA signal measured with respect to Vcm is launched down a 50Ω transmission line.
13.4 How Are Differential Transmission Lines Terminated?
169
From this we see that, although the traces were routed with a 50Ω odd-mode
impedance, the differential impedance is actually twice that (100Ω).
13.4 How Are Differential Transmission Lines Terminated?
From Figure 13.8 we see that a proper termination is formed by connecting a resistor equal to the transmission line impedance between each traced and a commonmode voltage. For instance, for a diff-pair having a 50Ω odd-mode impedance,
attaching to each of the traces 50Ω resistors connected to a supply voltage (such as
the 0.5V common mode from the previous section) would properly terminate the
line.
A second approach appears in Figure 13.9, which shows that a valid termination can be formed by placing resistance directly across the traces forming the
diff-pair. It is evident the value for a single resistor would be twice the value of RT,
and we have just seen that for proper termination RT should be made equal to the
odd-mode impedance. For instance, attaching a 100Ω resistor across the diff-pairs
will properly terminate them when they have an odd-mode impedance of 50Ω. This
makes intuitive sense: Because the differential impedance is equal to twice the oddmode impedance, a single resistor equal to the differential impedance should be
a proper differential termination. These two ideas are illustrated in Figure 13.10.
Figure 13.10(a) shows the common-mode termination, while Figure 13.10(b)
shows the differential termination technique. We can see in Figure 13.10(a) that if
the common-mode voltage is made zero, the two resistors are effectively connected
to ground. In fact, diff-pairs are sometimes terminated with a resistor equal to the
odd-mode impedance (Zoo) connected between each of the traces and ground. This
scheme is attractive for those drivers that require a connection to ground to establish a DC bias for proper operation, but such a termination is not truly differential.
Any difference in the voltage between the ground at the transmitter and the local
ground at the receiver where the termination resistors are connected will appear as
a voltage offset to the receiver, which can erode receiver noise margins.
Zoo = 50Ω
50Ω
0.75V
RI+
DO+
VDO+
+
5mA
0.5V
RT
0.5V
Vcm
+
0.5V
RT
DO−
0.25V
50Ω
RI−
Zoo = 50Ω
Figure 13.9 Output voltages measured with respect to ground.
170
Differential Signaling
Zoo
RI+
DO+
RT = Zoo
+
Vcm
−
RT = Zoo
DO−
RI−
Zoo
(a)
Zoo = 50Ω
RI+
DO+
+
RT = 2Zoo
−
DO−
RI−
Zoo = 50Ω
(b)
Figure 13.10 Diff-pair termination either (a) with a resistor to the common mode voltage or (b)
between traces.
13.4.1 How Should a Diff-Pair Be Terminated When the Propagation Is Not
Fully Odd Mode?
Assuming that signals propagate in the odd mode with no even-mode component
is an important caveat that is usually not valid in practice, especially when working with high-speed signals. We now turn our attention to describing a few of the
physical effects that can cause a differential signal to experience mode conversion
(changing from fully odd mode to having some even-mode component) [21, 22].
In the previous section we only had a single propagation mode (odd mode), so
a single resistor across the diff-pairs was sufficient to properly terminate the traces.
However, when both the even and odd modes are present, it is necessary to properly terminate both modes. Figure 13.11(a) shows how to do this.
Three resistors are necessary: one on each trace to a voltage source to terminate
the even mode (R1 in the figure) and one connecting the traces for the odd- mode
termination (R2). The voltage source can be a termination voltage or it can be set
to zero, which is the same as connecting the resistors to ground, as is shown in
Figure 13.11(a).
The values for the resistors [23] are given in:
13.4 How Are Differential Transmission Lines Terminated?
171
R1
Zoo
DO+
RI+
+
R2
−
DO−
RI−
Zoo
R1
(a)
Zoo
C1
RI+
DO+
R1
R2
Vtt
+
−
R1
DO−
Zoo
RI−
C1
(b)
Figure 13.11 (a) Termination scheme when the even and odd modes cannot be ignored. (b) By
connecting R1 to Vtt the line is simultaneously rebiased and terminated. Capacitor C1 allows the
transmitter and receiver to operate at different bias voltages.
R1 = Zoe
R2 =
2Zoe Zoo
Zoe − Zoo
(13.4)
(13.5)
Figure 13.11(b) shows how capacitors and a power supply can be used to
simultaneously rebias and terminate a diff-pair. The ability to both level shift and
terminate is useful when the receiver and transmitter common-mode voltages are
different. For instance, this technique is used when it is necessary to send signals
between logic families operating at different supply voltages. In this case Vtt is set to
the common-mode voltage required by the receiver and so puts the diff-amp at the
optimum bias point for its operation. The transmitter can be signaling at a voltage
significantly different from Vtt, but the capacitors only allow the signal amplitude
172
Differential Signaling
to pass and do not allow the receiver to be influenced by the transmitter’s different
common-mode voltage. This idea is explored in the Problems.
From the Problems we find that (13.4) and (13.5) can be checked by noting
that when the traces are very loosely coupled, R2 becomes infinite and R1 becomes
equal to Zo. This makes intuitive sense: Very loosely coupled lines can be treated as
isolated transmission lines and are properly terminated with a resistor equal to Zo.
Some field solvers directly report the even- and odd-mode impedances, which
makes using (13.4) and (13.5) easy. Alternatively, the even and odd impedances can
be calculated by using (9.4) and (9.6) from Chapter 9, as we do in this chapter’s
Problems when finding R1 and R2.
13.4.2 What Prevents Creating Odd-Mode Signals?
As we have seen, the diff-pair signals must be exactly out of phase for odd-mode
signaling. In fact, anything that causes the two signals to lose this phase relationship (even for a short distance) creates an even-mode voltage that, if large enough,
requires the three-resistor scheme shown in Figure 13.11 for proper termination.
Besides being a signal integrity problem, even-mode currents are a serious electromagnetic interference (EMI) concern that can hinder a product from passing radiated emission type testing [24].
Even mode signaling can arise when:
•
The diff-pair trace lengths are not exactly the same length.
•
The diff-pairs are routed in different dielectrics (microstrip versus stripline)
or on different routing layers.
•
The transmitter output drivers do not switch at precisely the same time.
•
The output drivers have different rise and fall times.
Layout techniques (presented in the next section) can help minimize length
matching problems, and proper design of the transmitter output driver (along with
the logic that drives it) can minimize the remaining two items.
Circuit board signal integrity engineers perform simulations or electrical measurements to determine how closely the driver output impedance and rise times are
matched, and the results from field solvers are used to determine if a significant
even-mode component is present in the diff-pairs. If so, it is necessary to use the
three-resistor scheme shown in Figure 13.11. By building detailed circuit models,
the effects that the imperfect driver and unequal trace lengths have on the overall
circuit operation can be determined.
13.5 How Are Differential Transmission Lines Created?
The two traces forming a diff-pair can either be routed side by side on the same
layer (edge-coupled) or on top of each other on separate layers (broadside-coupled).
Cross-sections of these two configurations (topologies) are shown in Figure 13.12.
They may be formed as tightly or loosely coupled pairs, but, as we will see, these
topologies have different electrical characteristics.
13.5 How Are Differential Transmission Lines Created?
173
s
s
W
b
h
(a)
(b)
(c)
Figure 13.12 (a, b) Edge-coupled and (c) broadside-coupled diff-pairs in cross-section.
13.5.1 Manufacturing Trade-Offs Between Loosely and Tightly Coupled Pairs
Loosely coupled pairs are attractive to fabricate because the alignment and spacing between traces and to the return planes is not as critical as with tightly coupled
traces. Also, the differential impedance of loosely coupled traces is not affected as
much by over etching. This means that it is easier for the manufacturer to obtain the
specified differential impedance even if the traces are not rectangular (see Chapter
8). Both of these factors reduce scrap and so makes manufacturing less costly.
Another consideration is that by making edge coupled traces loosely coupled
the individual routing layers are thinner than they would be if the traces were
tightly coupled. As illustrated in Table 13.1, this makes it possible to include more
routing layers for a given overall board thickness.
The microstrip traces are covered with 1 mil of solder mask having the same
dielectric constant as the laminate.
In Table 13.1 the tightly edge-coupled microstrip and stripline traces are defined as having a trace separation of (s) of equal to the trace width, while s is three
times the trace width for loosely coupled traces. This designation is arbitrary, but it
allows us to see how choosing the amount of coupling can affect the routing layer
thickness.
On FR4 separating the 4-mil-wide broadside traces by less than about 5 mils
requires impractical board dimensions, so for the broadside case tight coupling is
defined in Table 13.1 as 7 mils. This requires in a plane separation (b) of 31 mils
to obtain a 100Ω differential impedance. Loosely coupling the traces by increasing the trace separation to 20 mils results in a slight reduction in the overall layer
thickness.
Table 13.1 Approximate Values for h and
b (in Mils) for Tightly and Loosely Coupled
4-Mil-Wide, Half-Ounce 100Ω Diff-Pairs on
FR4 (εr = 4.2)
Topology
Tight
h or b
s
Loose
h or b
s
Microstrip 4
4
3
12
Stripline
42
4
11
12
Broadside
31
7
29
20
The separation between traces is s (in mils). Figure 13.12
defines h, b, and s.
174
Differential Signaling
13.5.2 Electrical Trade-Offs Between Loosely and Tightly Coupled Pairs
There are three electrical characteristics to consider when choosing between loosely
and tightly coupled traces: signal loss, EMI, and noise.
In general, tightly coupled diff-pairs have lower signal losses than loosely coupled pairs when fabricated on the same laminate and with the same trance width.
Losses at 1 GHz are shown in Table 13.2 when the dielectric thicknesses are as
illustrated in Table 13.1.
RF emissions will be reduced with tightly coupled traces [27, 28], so loosely
coupled diff-pairs should be avoided (especially for microstrip diff-pairs) in those
circumstances where EMI is a concern.
13.6 What Are Some Suggested Diff-Pair Layout and Routing Rules?
The electrical requirements for proper differential signaling are:
•
The differential impedance must be consistent and of the correct value.
•
The traces forming the diff-pairs must be the same length.
•
Noise must be equally coupled to both traces.
These three electrical rules are satisfied with the following rules:
1. Use the trace width and spacing specified by the circuit board manufacturer
to obtain the proper differential impedance.
2. Do not allow a signal trace to come between the two traces forming a
diff-pair.
3. Keep signal traces (and vias) much further from the diff-pair than the spacing between the pairs themselves.
4. Avoid routing over splits in the return path.
5. Route the diff-pair traces on the same layer.
6. Minimize layer hopping.
7. Use differential vias when layers must be changed.
8. Match the overall trace lengths.
9. Match the length on each layer.
10. Route traces passing through pin fields and near vias so that noise is equally
coupled on both pairs.
Table 13.2 Loss (in Decibels Per Meter)
Simulated [25, 26] at 1 GHz for Tightly and
Loosely Coupled 4-Mil-Wide 100Ω Diff-Pairs
on FR4 (εr = 4.2)
Topology
Tight
Loose
Microstrip
7.9
8.5
Stripline
8.3
8.4
Broadside
8.2
8.9
13.6 What Are Some Suggested Diff-Pair Layout and Routing Rules?
175
An explanation and example violations for each rule appear in the following
sections.
13.6.1 Obtaining the Proper Differential Impedance
Differential pairs are so common in modern designs that nearly all circuit board
fabrication shops now have experience in creating them. This is especially true
for edge-coupled 100Ω diff-pairs that are wider than the minimum allowed trace
width. For instance, most shops can reliably create an edge coupled 100Ω diff-pair
using 8-mil-wide (or wider) traces. High-end shops can reliably form 100Ω diffpairs with traces less than 5 mils wide.
It is important not to overspecify when discussing the requirements with the
circuit board shop. For instance, it is proper to specify that the traces are to be a
100Ω differential pair using traces 8 mils in width and that the traces are to be
spaced as close together as possible. This gives the shop freedom to adjust the trace
width (bias the trace), spacing, and laminate thickness based on their experience
manufacturing similar boards. However, it is undesirable to tell the shop that the
traces are to be 8 mils wide and have a 50Ω odd-mode impedance when spaced
8 mils apart. Although technically correct, this terminology not widely used by
circuit board manufacturers and is so restrictive that it might increase the board
manufacturing cost.
The second aspect of maintaining proper differential impedance involves trace
routing. Once the manufacturer has specified the trace separation necessary to obtain the desired differential impedance, make sure the layout designer adheres to
this everywhere.
One example of how this rule is violated is shown in Figure 13.13.
In this example a diff-pair jogs around a via, which for a short distance increase
the trace-to-trace spacing. This encourages the traces to couple to the via rather
than to each other, altering the impedance.
The third aspect of maintaining proper differential impedance involves trace
width. It is sometimes necessary to intentionally make a trace narrow so it can be
routed through a connector or BGA pin field. This is especially likely with long
traces because they are often made very wide to reduce signal loss. Such wide
DO+
DO−
Figure 13.13 Top view of diff-pair traces jogging around a via and its antipad (shown as a broken
circle). This routing increases the differential impedance.
176
Differential Signaling
traces must be reduced in width (necked down) to fit through the pin field, which
increases the trace impedance through that region.
In practice the communication link may be robust enough to operate satisfactorily when the signal is exposed to any of these types of discontinuities. This is
particularly true if the data rate is low enough to make the period of the discontinuity small compared to the signal bit time. However, even a low data rate link can
fail if many of these discontinuities are simultaneously present.
13.6.2 Signal Trace Between Diff-Pairs
When routing edge-coupled diff-pairs, insure that there are no signal traces routed
between the traces forming the diff-pair. It is not usually possible to violate this rule
when the traces are tightly coupled, but violations can arise when the traces are
widely separated (loosely coupled).
Violation of this rule usually comes about when space between the diff-pairs
is large enough for the autorouter to place a signal trace in the area between the
diff-pairs. Autorouting software is a tool often used by layout designers because
it speeds the layout process by working from a list of nets to automatically place
and connect traces to components. If there is enough space, the autorouter may
route signals between the traces forming the loosely coupled diff-pair unless it is
purposely restricted from doing so. This type of error can be hard to find on dense
boards because the errant trace (or traces) may only run for a short distance.
13.6.3 Keeping Signal Traces Far Away from Diff-Pairs
The top view of either a stripline or microstrip edge coupled diff-pair is shown in
Figure 13.14. Traces A1 and A2 are aggressors routed alongside the diff-pair (DO+
and DO−). It is apparent that trace A1 will preferentially couple onto DO+ and
will couple much less to DO−. This asymmetric coupling changes the impedance of
DO+ and makes the noise common mode, which will not be rejected by the receiver.
For these reasons the separation from the aggressor to the diff-pair (sn) must
always be much greater than the separation between the traces forming the diffpair (sd). This rule applies to either broadside or edge-coupled pairs.
A1
sn
DO+
DO−
sd
A2
Figure 13.14 Top view showing four traces. Prevent common mode noise by making the spacing
between the diff-pair (sd) smaller than the spacing to other traces (sn).
13.6 What Are Some Suggested Diff-Pair Layout and Routing Rules?
177
13.6.4 Avoid Return Path Splits
The most obvious application of this rule is when a diff-pair crosses over a split in
a power or ground plane.
Splits are cuts in the plane that create isolated regions (“power islands”). They
are often used to divide a power plane so that multiple voltages can be located on
the same plane. For instance, the ground plane is sometimes split to isolate an analog ground from a digital ground.
High-speed, loosely coupled differential signals are more likely to have difficulty crossing a split. In contrast, slower-speed differential signals, especially when
they are tightly coupled and are high-impedance, are better able to tolerate splits.
For these types of traces, the return current is mostly in the diff-pair traces and not
in the ground plane, so a void in the ground plane may not significantly alter the
transmission line characteristics.
A more subtle version of the same effect is illustrated in Figure 13.15.
Figure 13.15 shows a trace crossing a void in the ground plane caused by a
large antipad. Antipads (first introduced in Chapter 5) are holes in the power or
ground planes that allow a via to pass through without making connection. Antipads are sometimes made with large diameters to reduce the coupling capacitance
from the via to the plane. This is especially common on boards having many layers
(including back planes) because reducing the parasitic capacitance of each layer by
even a small amount can significantly reduce the total parasitic capacitance.
The difficulty illustrated in Figure 13.15 is that the local ground plane has been
removed in the region where the trace crosses the antipad. This increases the trace
inductance and lowers its capacitance, raising the trace impedance in that region. A
trace routed through a deep pin field may cross over many of these discontinuities,
which can result in severe signal degradation.
13.6.5 Route on Same Layers
The two traces forming an edge coupled diff-pair should always be routed on the
same layer. Otherwise, the delay, impedance, and noise coupling are very unlikely
to be identical.
This rule is sometimes violated when the two capacitors connecting a serial
transmitter or receiver—generally, a serializer/deserializer (SERDES)—to a diffpair are placed on opposite sides of a circuit board. Often this technique allows the
capacitors to be placed closer to the SERDES, but doing so makes the noise and
impedance on the two traces unequal. It also usually results in different via lengths,
making the overall length of the two traces unequal.
Figure 13.15 A trace crossing over a large antipad (dotted). The trace impedance changes in the
region where it crosses the split.
178
Differential Signaling
13.6.6 Minimize Layer Hopping
Vias represent an impedance discontinuity unless their impedance is intentionally
designed to match the trace impedance (described in Chapter 14). This gives rise to
reflections and at high speeds can significantly impair the ability of a serial interconnect from operating properly.
13.6.7 Use Differential Vias
A pair of vias can be designed to have differential impedance equal to the differential impedance of the diff-pair traces. As described in Chapter 14, designing
these structures requires a 3D field solver to model the effects of via size, antipad
diameter, pad diameter, and laminate thickness on the differential impedance [29].
Such via structures greatly reduce (and may eliminate) the impedance discontinuity
presented by vias not as carefully designed, thereby reducing signal reflections and
allowing the signal to operate at very high speeds.
13.6.8 Match Overall Trace Lengths
Violations of this rule can be subtle and detecting them often requires a detailed
understanding of the routed path.
For instance, right-angle multirow connectors have different length pins in the
upper and lower rows. If these two rows are used to form a diff-pair, the length
skew occurs on each of the connectors forming a mated pair, doubling the difference. For this reason routing diff-pairs through these connectors should be done
with pins on the same row. This length mismatch can be present when routing
either edge-coupled or broadside traces.
This rule can be unintentionally broken when routing broadside diff-pairs. Because they are routed on different layers, vias or interlayer connections made by
connector pins moving signals to the board’s outer surface naturally will be of different lengths [13]. When two connectors are used (as in a midplane or daughter
card situation) this inevitable length skew can sometimes be corrected by swapping
the layers on the two cards. If the two cards have the same stackup thickness the
via (or pin) lengths become matched, eliminating the skew.
A third subtle way in which this rule is violated is shown in Figure 13.16 where
a trace includes one or more bends as it changes direction.
Figure 13.16(a) shows how the bends cause a difference in trace length. In
high-performance signaling back-to-back bends can be used to readjust the overall
length.
13.6.9 Match Lengths on Each Layer
Besides ensuring that the total length of trace from the transmitter to receiver is
the same for the two traces forming a diff-pair, it is important that the lengths are
matched on each of the layers on which they are routed. For instance, a diff-pair
may exit a BGA pin field on the surface, using microstrip, and then transition with
a pair of vias to a stripline layer. The lengths of the two microstrip traces should be
13.6 What Are Some Suggested Diff-Pair Layout and Routing Rules?
179
DO+ DO−
(a)
DO+
DO−
(b)
Figure 13.16 (a) Bends in diff-pairs cause DO+ to be longer than DO−. (b) Back-to-back bends can
readjust this skew.
matched, and the lengths of the two stripline traces should be matched. This ensures
that the delay and impedance of each segment is the same.
13.6.10 Ensure That Noise Is Equally Coupled to Both Traces
Subtle mechanisms can cause unequal noise coupling on the diff-pair traces. An example showing the effects on a broadside-coupled pair placed between two return
planes appears in Figure 13.17.
Noise present on plane L1 will preferentially couple onto the L2 trace, and
noise on L4 will couple onto L3. The traces forming the diff-pair are subjected to
different noise unless the planes have a low impedance between them and have the
same noise voltages. If both L1 and L4 are ground planes, placing numerous vias
between the planes (plane stitching) can lower the impedance between the planes.
Alternatively, if one is a power plane, decoupling capacitors can be used to provide
an AC connection between the planes. In this way the same noise appears on both
planes, making the noise coupled from L1 to L2 identical to the noise coupled from
L4 to L3.
Another effect that can occur with edge coupled diff-pairs (either stripline or
microstrip, but not with broadside pairs) is shown in Figure 13.18.
180
Differential Signaling
L1
h
L2
s
L3
h
L4
Figure 13.17 Side view of broadside coupled diff-pair formed on routing layers L2 and L3. Noise
on the L1 plane will preferentially couple onto L2, and noise on plane L4 will more readily couple
onto L3.
A4+
A4−
A3+
A3−
A2+
A2−
A1+
A1−
S1+
S1−
Figure 13.18 Top view of a diff-pair routing through a pin field. The noise environment for S1+ is
not the same as for S1−. (From: [13]. Used with permission.)
In Figure 13.18 an edge-coupled diff-pair is routed through a pin field, such as
a dense connector field or a BGA. The circles represent vias to which the connector
is inserted, although this also occurs with vias connecting to surface mount lands.
Ground connections are represented by dark circles, and the remaining circles are
differential signals.
The signal connected to the S1+ pin experiences a very different noise environment than its complement (S1−). Signal S1+ is routed between rows of differentially
switching signals, making the net transfer of noise zero. However, the S1− signal is
routed between ground and the switching signals. In this case the switching noise
does not cancel and noise from the A1− through A4− pins accumulates on the
S1− signal. Since they are on top of one another, broadside pairs are immune from
13.7
Main Points
181
this as they will experience the same pin-to-trace coupling no matter how the are
routed.
13.7
Main Points
•
Differential signaling uses a differential amplifier to reject noise present on
both its inputs, making it important to couple noise equally to both inputs.
•
The differential impedance of a differential transmission line is twice the oddmode impedance.
•
A single resistor placed across the diff-pair is a proper termination for diffpairs with negligible even-mode energy.
•
A three-resistor termination scheme is used when the even-mode component
is too large to be ignored.
•
Even-mode behavior is caused by poor routing practices or poor differential
driver characteristics.
•
Diff-pairs can be formed as edge or broadside coupled, and as either loosely
or tightly coupled.
•
Loosely coupled diff-pairs are easy to manufacture but tightly coupled traces
may be electrically necessary.
•
Special test equipment is necessary to analyze differential signals in detail.
•
Eye diagrams clearly show the effects of noise and interference present on
the channel.
Problems
Answers to these problems are available on the Artech House Web site at http://
www.artechhouse.com/static/reslib/thierauf/thierauf1.html.
13.1 A field solver produced the following output file for a diff-pair.
L MATRIX [H/INCH]
1.052E-8 2.040E-9
2.040E-9 1.052E-8
C MATRIX [F/INCH]
2.835E-12 −5.490E-13
−5.490E-13 2.835E-12
Find the values for termination resistors R1 and R2 in Figure 13.11.
13.2 A diff-pair is described by the following SPICE transmission line model:
+ Lo = 8.92E-09
+ 1.32E-09 8.92E-09
+ Co = 2.50E-12
+ -2.34E-13 2.50E-12
What is the differential impedance?
182
Differential Signaling
13.3 The traces described in Problem 13.2 show a difference in the odd- and
even-mode impedances of more than 25%. Is it necessary to use the two
resistor termination shown in Figure 13.11, or will a single resistor across
the diff-pair [as shown in Figure 13.10(b)] be sufficient?
13.4 It has been decided that a single termination resistor, placed across the diffpair (as shown in Figure 13.10) will be used to terminate the diff-pair described in Problem 13.2. What should its value be?
13.5 It has been decided that resistors to ground from each trace in the diff-pair
will be used to terminate the traces described in Problem 13.2. What values
should these resistors have?
13.6 A loosely coupled diff-pair is described by the following SPICE transmission
line model:
+ Lo = 8.47E-09
+ 2.30E-10 8.47E-09
+ Co = 3.08E-12
+ -2.43E-14 3.08E-12
What are the proper termination values if the termination appearing in Figure 13.11 is used?
13.7 A clock oscillator has a differential output that swings from 0.8V to 1.8V
peak-to-peak (p/p). It is necessary to connect this to an ASIC differential receiver that requires the signal to have a 1V p/p swing centered on
1.65V. How should the diff-pairs connecting the oscillator to the ASIC be
terminated?
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
“High-Speed Transceiver Logic (HSTL) A 1.5V Output Buffer Supply Voltage Based Interface Standard for Digital Integrated Circuits,” EIA/JESD8-6, Electronics Industries Association, August 1995.
“JDEC STANDARD: Stub Series Terminated Logic for 1.8 V (SSTL_18),” JESD8-15A,
JEDEC Solid State Technology Association, September 2003.
National Semiconductor, LVDS Owner’s Manual, 4th ed., Santa Clara, CA: National Semiconductor Corporation, 2008.
Infiniband™ Trade Association, “Infiniband ™ Architecture Specification, Vol. 2, Release
1.1,” November 2002.
Fuller, S., RapidIO: The Embedded System Interconnect, New York: John Wiley & Sons,
2005.
Hug, S., and J. Goldie, “An Overview of LVDS Technology,” National Semiconductor Applications Note 971, National Semiconductor Corporation, 1998.
Tektronix Incorporated, Beverton, OR.
Lauterbach, M., “Getting More Out of Eye Diagrams,” IEEE Spectrum, Vol. 34, No. 3,
March 1997.
Bateman, A., Digital Communications, Reading, MA: Addison-Wesley Longman, 1998.
Dally, W. J., and J. W. Poulton, Digital Systems Engineering, Cambridge, U.K.: Cambridge
University Press, 1998.
Couch, L. W., II, Digital and Analog Communication Systems, 5th ed., Upper Saddle River,
NJ: Prentice-Hall, 1997.
Problems
[12]
[13]
[14]
[15]
[17]
[16]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
183
Bissell, C. C., and D. A. Chapman, Digital Signal Transmission, Cambridge, U.K.: Cambridge University Press, 1996.
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Horowitz, M., et al., “High-Speed Electrical Signaling: Overview and Limitations,” IEEE
Micro, Vol. 18, No. 1, January/February 1998, pp. 12–24.
Proakis, J. G., Digital Communications, New York: McGraw-Hill, 2000.
Terman, F. E., Radio Engineers Handbook, New York: McGraw-Hill, 1943.
Haykin, S., and B. Van Veen, Signals and Systems, 2nd ed., New York: John Wiley & Sons,
2002.
Dally, W. J., Poulton, J., “Transmitter Equalization for 4-Gbps Signaling,” IEEE Micro,
Vol. 17, No. 1, January/February 1997.
Turudic, A., et al., “Pre-Emphasis and Equalization Parameter Optimization with Fast,
Worst-Case/Multibillion-Bit Verification,” Design Con 2007, Santa Clara, CA, January
29–February 1, 2007.
Chandrakasan, A., et al., Design of High-Performance Microprocessor Circuits, New
York: IEEE Press, 2001.
Young, B., Digital Signal Integrity, Upper Saddle River, NJ: Prentice-Hall, 2001.
Simbeor Applications Note #2009_1, “Practical Notes on Mixed-Mode Transformations
in Differential Interconnects,” Simberian Corporation, Seattle, WA, January 2009.
Marx, K. D., “Propagation Modes, Equivalent Circuits, and Characteristic Terminations
for Multiconductor Transmission Lines with Inhomogeneous Dielectrics,” IEEE Transactions on Microwave Theory and Techniques, Vol. MTT-21, No. 7, July 1973, pp. 450–457.
Paul, C. R., Introduction to Electromagnetic Compatibility, New York: John Wiley &
Sons, 1992.
Polar Si9000 Field Solver, Polar Instruments, Inc. Beaverton, OR.
Djordjevic, A. R., et al., LINPAR for Windows, Norwood, MA: Artech House, 1999.
Ott, H. W., Noise Reduction Techniques in Electronic Systems, 2nd ed., New York: John
Wiley & Sons, 1988.
Rainal, A. J., “Transmission Properties of Balanced Interconnections,” IEEE Transactions
on Components, Hybrids and Manufacturing Technology, Vol. 16, No. 1, February 1993,
pp. 137–145.
Simbeor Application Note #2007_07, “Modeling Differential Via-Holes Without Stitching
Vias,” Simberian Corporation, Seattle, WA, October 2007.
CH A P T E R 14
Trace and Via Artwork Considerations for
Signal Integrity
14.1
Introduction
Practical considerations can cause the actual circuit board artwork to differ from
the design that was simulated. Some of these things (such as routing close to the
board edge or the way in which signals are length matched with serpentine traces)
are visible in the artwork, but other choices (such as the sizes of the vias and antipads) are not as immediately obvious. The signal integrity engineer who is aware
of these practical concerns can deliberately look for these structures as part of the
layout artwork checking process and update the simulation model as necessary.
14.2
Mitered Corners
Nearly all traces will have one or more bends as they change direction or jog around
obstacles. As shown in Figure 14.1, a sharp right angle bend increases the trace area
in the vicinity of the bend. This adds a small amount of excess capacitance [1, 2]
and excess inductance [3, 4] to the trace.
The excess inductance and capacitance are very small, but alter the impedance
at the bend, which can cause reflections at high frequencies. Usually the excess
capacitance is the dominate effect and is typically in the hundreds of femto farad
range [5].
Mitering (or chamfering) is commonly used to reduce the excess capacitance
and is shown in Figure 14.2. Mitering is not necessary until the signals have rise
times in the subnanosecond range [5], but it can reduce reflections for those traces carrying frequencies (including harmonics) greater than several gigahertz [6].
Notches [6, 7] are sometimes used at microwave frequencies, but these exotic approaches are generally not seen on digital circuit boards.
The theoretical optimum miter occurs when the excess capacitance has been
reduced to where the capacitive and inductive effects are equal [8], but in practice
the actual angle and length of the miter (m) are not usually critical. As shown in
Figure 14.2, an angle of 45° is commonly used, and the miter length can range from
185
186
Trace and Via Artwork Considerations for Signal Integrity
W
Figure 14.1 A trace of width W making a right-angled turn. The extra trace area (shown in white)
adds capacitance and inductance, which can cause reflections at high frequency.
m = 0.8W = 6.4 mils
W = 8 mils
m = 2.25W = 10 mils
W = 8 mils
Figure 14.2 Mitering the trace outside edge at 45° to reduce excess capacitance. Optimal miter
length (m) is between 0.8W and 1.25W. Dimensions are shown for an 8-mil (0.21-mm)-wide trace.
0.8W [9] to 1.25W [6]. For instance, m for an 8-mil (0.21-mm)-wide trace would
be between 6.4 and 10 mils (0.17 mm and 0.25 mm) at an angle of 45°. In practice
14.3 Routing Near the Board Edge
187
the layout designer often selects these values such that the manufacturer’s minimum
trace width rules are not violated, especially on narrow traces.
14.3 Routing Near the Board Edge
Some of the electric and magnetic field lines escape traces which are routed close to
the circuit board edge, increasing the trace inductance and decreasing its capacitance.
These changes reduce the delay but increase the trace impedance and conductor losses. The effects are most pronounced with wide, high-impedance traces because they are further from the return planes than narrow, low-impedance traces.
Losses increase because the return current cannot spread as far, increasing return
path resistance.
A top view of two 10-mil-wide, 65Ω stripline traces running parallel to the
edge of the board is shown in Figure 14.3(a). One trace is 5 mils (0.13 mm) from
5 mils
100 mils
L = 11.23 nH/inch
C = 2.68 pf/inch
Z = 65Ω
tpd = 174 ps/inch
L = 12.31 nH/inch
C = 2.24 pf/inch
Z = 74Ω
tpd = 169 ps/inch
(a)
5 mils
100 mils
20 mils
(b)
Figure 14.3 10-mil (0.25-mm)-wide stripline routed 5 mils (0.13 mm) from the board edge has
different electrical properties than identical stripline routed much further in (a). (b) Side view shows
the traces are 20 mils (0.5 mm) from either plane.
188
Trace and Via Artwork Considerations for Signal Integrity
the edge, while the other, identical trace is 100 mils (2.5 mm) from the other edge.
A side view appears in Figure 14.3(b).
The inductance of the stripline routed close to the board edge is about 10%
higher and the capacitance is over 15% lower than the other stripline. This causes
the impedance to grow by nearly 14% and the delay to fall by nearly 3%. Although
not shown, the impedance and delay of a 4-mil-wide, 50Ω stripline (which is significantly closer to the return plane then the traces in Figure 14.3) only change by
about 2% under these same conditions.
Microstrip behaves in the same general way as stripline, with the biggest effects
seen on wide, high-impedance traces. Often 2D field solvers tacitly assume the return plane is very much wider than the trace. Figure 14.3 shows that circuit models
created with this assumption will not represent the characteristics of the traces
routed close to the edge. This can be a significant problem when signaling at high
speeds, particularly if the signaling is occurring with differential pairs.
14.4
Serpentine Traces
Often high-performance integrated circuits include delay locked loops (DLLs) or
similar circuits to digitally adjust the timing between signals [11, 12]. When these
circuits are not present or when it is necessary match the arrival of individual signals within a bus, a long trace can be used to add a known delay based on the transmission lines time of flight (tpd). Because tpd is typically hundreds of picoseconds
per inch, adjusting timing usually requires several inches of trace.
These delay lines are created by folding the trace back and forth to create a
serpentine trace. They are also referred to as meander or trombone traces. A fivesegment serpentine is diagrammed in Figure 14.4.
Each of the five segments has a delay equal to tpd, making the expected total
delay 5 tpd plus the delays of the short vertical connecting straps. However, the
actual delay will be less than this unless care is taken with the layout.
1
Kb1,2
2
3
Kb1,5
4
5
tpd
Figure 14.4 Five-segment serpentine delay line. Each long segment has a delay of tpd.
14.4
Serpentine Traces
189
Near-end crosstalk (discussed in Chapter 10) between traces 1 and 5 (Kb1,5)
causes trace 1 to induce a small voltage on trace 5 and a larger voltage on trace
2 from (Kb1,2). This coupling is shown as occurring on trace 3 because here the
vertical connecting strap is assumed to be short enough to be ignored. The Kb1,5
signal arrives at the serpentine output tpd seconds after traveling the length of
trace 5, and the stronger Kb1,2 signal arrives 2 tpd seconds later, after traveling
the entire length of traces 3 and 4. This voltage adds to the crosstalk already present, causing the output to increase. The output reaches the full value at 5 tpd with
the arrival of the launched signal. A ladder-like waveform [13] has been created at
the serpentine output, as illustrated in Figure 14.5(a). The expected waveform is
shown in Figure 14.5(b).
Laddering can occur with either stripline or microstrip because each are susceptible to near end crosstalk.
V
tpd
2 tpd
2 tpd
(a)
V
5 tpd
Voltage
Time
(b)
Figure 14.5 (a) Ladder-like waveform created by crosstalk appearing at the output of a five-segment serpentine delay line contrasts with (b) the expected waveform.
190
Trace and Via Artwork Considerations for Signal Integrity
14.4.1 How Are Serpentine Delay Lines Modeled?
Multiconductor lossy circuit models of the serpentine can be created with a 2D
field solver. However, this technique is only an approximation because it does not
account for the coupling occurring from a horizontal segment to a vertical strap.
A 3D solver that includes the entire serpentine structure is necessary to account for
these effects.
Once created, the models are used in a circuit simulator with models of the
actual driver and receiver to determine if a given spacing or number of segments
will create laddering severe enough to falsely trigger the receiver. A multiconductor
transmission line model represents the long horizontal segments, and the short vertical straps are modeled with direct connections. If they are long enough, transmission lines can be used for these straps. This technique is illustrated in Figure 14.6.
14.4.2 Effects of Corners
The serpentine corners (bends) have less delay than an equivalent amount of straight
trace and so are said to speed up the signal. Mitering the corners can help correct
this timing uncertainty [14].
Excess capacitance causes the impedance to be lower in the corners than the
segments. This can be significant if there are many corners or when signaling at
high speeds.
Corner effects and the coupling from the vertical segments to the horizontal
straps are not properly accounted for in circuit models created by a 2D field solver.
However, a 3D solver will capture them [15].
14.4.3 How Long Should the Segments Be?
Because the maximum coupled voltage grows with the number of segments in the
serpentine [16], when laying out a serpentine, it is best to use a fewer number of
long segments instead of a greater number of short ones [15]. Fewer segments also
mean fewer corners and less uncertainty in the timing and impedance.
For these reasons the segments should be long (typically greater than the signal rise time) and few in number. Also, because crosstalk increases as the traces
Zero delay
vertical
straps
tpd
1
Zo
5
Lossy
transmission
line model
Delay = 5 tpd
Figure 14.6 A multiconductor transmission line modeling a five-segment serpentine delay line. The
vertical straps can be wires as shown or transmission lines.
14.5
Vias
191
are tightly packed together, laddering can be reduced by increasing the separation
between segments.
14.4.4 Guard Traces
Placing a properly grounded guard trace between segments in the serpentine can
reduce or eliminate laddering since this reduces crosstalk between the segments.
Chapter 10 describes how to properly ground a guard trace. Adding them slightly
reduces the serpentine delay [13, 17], but makes the delay more uniform across a
broad range of frequencies [17].
14.5
Vias
A 3D rendering of a through hole via extending through a plane in a circuit board
is shown in Figure 14.7.
On each end the via, a circular ring of metal (“the pad”) connects to the signal
trace, and a hole (the antipad) in the plane permits the conducting barrel to pass
through without making a connection.
14.5.1 Via Circuit Model
Figure 14.8 shows how capacitance is present from the via barrel and the pads to
the plane. All of these capacitors are summed into a single capacitance, Ct. An inductor (Lt) models the inductance of the barrel and its return path.
Pad
Trace
Barrel
Antipad
Plane
Figure 14.7 A through hole via passing through a plane. The antipad prevents the barrel from connecting the plane. A pad connects the signal trace to the barrel. (Figure generated by the Simbeor
field solver [18].)
192
Trace and Via Artwork Considerations for Signal Integrity
Figure 14.8 Side view of a through hole via passing through a solid plane showing the parasitic
capacitance. Not shown is the parasitic inductance, which is also present.
A simple PI circuit, shown in Figure 14.9, can be used to model the impedance
and delay through the via when the frequency content is not too high.
The PI is formed by using the total inductance (Lt) and applying half of the
total capacitance (Ct) to each leg. This model is only valid when the voltage drop
across Lt and the current through the capacitors is small compared to the current
flowing through the via [19]. Typically this means the model is most accurate when
the signals have rise times of about 100 pS or longer [20]. At higher frequencies a
chain of these models, or a transmission line, may be used. For best results at still
higher frequencies, an S-parameter-type model can be used.
14.5.2 What Factors Determine Via Capacitance?
The total capacitance is determined by the diameter of the barrel and pads, by the
via length, and by the antipad diameter (and thickness and number of planes that
the barrel passes through) [21, 22].
The pad capacitance increases with the area of the pad, while the via barrel
capacitance increases as the square root of the length of the via [21]. These things
lower the via’s impedance. However, increasing the antipad size increases the distance from the barrel to the ground plane, which is equivalent to increasing the
plate separation in a parallel plate capacitor. This reduces the capacitance of the
barrel and tends to raise the via impedance.
Depending on the board thickness and the number of planes, a 10-mil (0.25mm) diameter through-hole via with a 30 mil (0.75 mm) antipad and 22-mil
Lt
Ct
2
Ct
2
Figure 14.9 This PI circuit is a good low-frequency circuit model of a via.
14.5
Vias
193
(0.55-mm) pad has a capacitance from between roughly 500 fF when the via is 60
mils (1.5 mm) long passing through five planes to 1.1 pF when it is 180 mils (4.5
mm) long [23].
14.5.3 What Factors Determine Via Inductance?
The return current will always find a path having the lowest impedance, and this
path determines the via’s total loop inductance (Lt).
The return current will preferentially flow through a nearby return via since
the small loop area has a low inductance. However, in the absence of a nearby
return via, the return current will seek alternate paths to lower the impedance.
This extended path is not accounted for in simple hand formulas that calculate via
inductance.
Additionally, as the impedance of the via path grows, to minimize the overall
return path impedance increasingly more of the return current flows through the
capacitance formed between the power and ground planes. This analysis is best
done by using a 3D field solver that includes the layer to layer capacitance and the
actual location of any return path vias or decoupling capacitors.
A 10-mil (0.25-mm) diameter through-hole via, with a 30-mil (0.75-mm) antipad and 22-mil (0.55-mm) pad and no local vias to act as return paths, has an
inductance varying from roughly 800 pH when it passes through five planes and is
60-mils (1.5 mm) long to 3 nH when it is 180-mil (4.5 mm) long [23].
14.5.4 What Are the Effects of Different Antipad Diameters?
A capacitor is formed each time the via passes through a plane. This is why a long
via with small antipads passing through many plane layers has a higher capacitance
and lower impedance than a short via with large antipads passing through a few
planes.
The effect of antipad size is illustrated in the time domain reflectometry (TDR)
plot appearing in Figure 14.10. The TDR shows the impedance in the vertical axis
against the distance in the horizontal direction. The circuit board used for this measurement has 20-mil (0.5-mm) diameter vias with 40-mil (1-mm) pads. The board
is 18 layers, 10 of which are power or ground planes.
From Figure 14.10 we see that in this one example that increasing the size of
the antipad from 45 mils (1.13 mm) to 60 mils (1.5 mm) increases the vias’ impedance from 32Ω to 43Ω.
14.5.5 What Are Nonfunctional Pads?
Especially in thick boards, manufacturers generally place pads on a via at each routing layer, even if a trace is not connecting to the via on that layer.
These nonfunctional pads are used to help relive thermal stresses placed on the
via during assembly operations [25, 26]. The stresses can cause via cracking when
the board is heated during soldering. Vias that are four or more times longer than
their diameter are most susceptible to cracking, especially when lead free soldering
is employed because of the high temperatures necessary for that process [25].
194
Trace and Via Artwork Considerations for Signal Integrity
75
Impedance (Ω)
High impedance
caused by probe
inductance
40 mil pad with 60 mil antipad
50
43Ω
32Ω
40 mil pad with 45 mil antipad
25
200 ps
400 ps
600 ps
Figure 14.10 TDR showing that a via with large antipad has higher impedance than one with a
small antipad. (From: [24]. Used with permission.)
The added capacitance of these pads lower the via impedance, which can cause
significant reflections when signaling at high speeds. The circuit board manufacturer can be requested to remove nonfunctional pads from those nets carrying high
data rate signals, but they should confirm that the board reliability will not be unduly impacted. Removal is done by the circuit board manufacturer and not in the
artwork by the layout designer.
14.5.6 Can the Via Impedance Be Adjusted?
The impedance of a signal carrying via can be controlled by the careful placement of
return path vias near the signal carrying via [27, 28]. This technique is of most use
when signaling at high data rates and requires the use of a 3D field solver to adjust
the placement of the vias and to select the pad and antipad sizes.
14.5.7 What Are Differential Vias?
A typical via presents a problem when differential signals change layers because the
via pair will not have the same differential impedance as the traces.
To correct this, a pair of vias can be designed for a specific differential impedance [29, 30] by adjusting their spacing and dimensions and occasionally even the
antipad shape. This is best done with a 3D field solver.
An example of a differential via appearing in a 3D field solver is shown in
Figure 14.11.
Coupling between the vias occurs in the region between the planes, so no other
structures (such as vias, microvias, traces, or nonplated holes) can be placed between the via pair. It is also important that other vias (including nonplated holes
14.6
Main Points
195
Figure 14.11 A pair of vias forming a differential pair passing through two planes. Ground stitching
vias are visible on either side. The rectangular tabs coming from the vias tops are microstrip traces.
(Figure generated by the Simbeor 3D solver [18].)
used by metal screws to secure the circuit board) not be placed close to either of the
vias forming the pair.
14.6
Main Points
•
Mitered corners can be used on traces to remove the excess capacitance and
inductance caused by bends.
•
The degree of chamfering usually is not critical.
•
Chamfering is important when signaling at high-speeds.
•
Routing traces near the board’s edge changes its electrical characteristics and
can increase EMI.
•
Crosstalk must be minimized when using serpentine traces to adjust timing.
•
It is better to use a few long serpentine elements than many short ones.
•
The via and antipad sizes strongly determine the vias electrical characteristics.
•
Nonfunctional pads help secure the vias in a stackup, but are not visible in
the artwork and can be removed on high-speed nets.
References
[1]
[2]
Silvester, P., and P. Benedek, “Microstrip Discontinuity Capacitances for Right-Angle
Bends, T Junctions, and Crossings,” IEEE Transactions on Microwave Theory and Techniques, Vol. MTT-21, No. 5, May 1973, pp. 341–346.
Silvester, P., and P. Benedek, “Corrections to Microstrip Discontinuity Capacitances for
Right-Angle Bends, T Junctions, and Crossings,” IEEE Transactions on Microwave Theory
and Techniques, Vol. MTT-23, No. 5, May 1975, p. 456.
196
Trace and Via Artwork Considerations for Signal Integrity
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
Thompson, A. F., and A. Gopinath, “Calculation of Microstrip Discontinuity Inductances,” IEEE Transactions on Microwave Theory and Techniques, Vol. MTT-23, No. 8,
May 1975, pp. 648–655.
Harms, P. H., and R. Mittra, “Equivalent Circuits for Multiconductor Microstrip Bend
Discontinuities,” IEEE Transactions on Microwave Theory and Techniques, Vol. MTT-41,
No. 1, January 1993, pp. 62–69.
Rainal, A. J., “Reflections from Bends in a Printed Conductor,” IEEE Transactions on
Components, Hybrids and Manufacturing Technology, Vol. 13, No. 2, June 1990, pp.
407–413.
Simbeor Application Note #2008_05, “Minimal-Reflection Bends in Micro-Strip Lines,”
Simberian Corporation, Seattle, WA, September 2008.
Widell, B. C., Transmission Line Design Handbook, Norwood, MA: Artech House, 1991.
Douville, R. J. P., and D. S. James, “Experimental Study of Symmetric Microstrip Bends
and Their Compensation,” IEEE Transactions on Microwave Theory and Techniques, Vol.
MTT-26, No. 3, March 1978, pp. 175–182.
Chadha, R., and K. C. Gupta, “Compensation of Discontinuities in Planar Transmission
Lines,” IEEE Transactions on Microwave Theory and Techniques, Vol. MTT-30, No. 12,
December 1982, pp. 2151–2156.
Archambeault, B., “Effects of Routing High-Speed Traces Close to the PCB Edge,” Printed
Circuit Design & Manufacture, January 2008.
Dally, W. J., and J. W. Poulton, Digital Systems Engineering, Cambridge, U.K.: Cambridge
University Press, 1998.
Chandrakasan, A., et al., Design of High-Performance Microprocessor Circuits, New
York: IEEE Press, 2001.
Wu, R., and F. Chao, “Laddering Wave in Serpentine Delay Line,” IEEE Transactions on
Components, Packaging, and Manufacturing Technology, Part B, Vol. 18, No. 4, November 1995, pp. 644–650.
Rubin, B. J., and B. Singh, “Study of Meander Line Delay in Circuit Boards,” IEEE
Transactions on Microwave Theory and Techniques, Vol. 48, No. 9, September 2000, pp.
1452–1460.
Orhanovic, N., et al., “Characterization of Microstrip Meanders in PCB Interconnects,”
Proceedings 50th IEEE Electronic Components and Technology Conference, Las Vegas,
NV, May 21–24, 2000, pp. 508–512.
Shiue, G., et al., “Improvements of Time-Domain Transmission Waveform in Serpentine
Delay Line with Guard Traces,” IEEE International Symposium on Electromagnetic Compatibility, EMC 2007, Honolulu, HI, July 9–13, 2007, pp. 1–5.
Nara, S., and K. Koshiji, “Study on Delay Time Characteristics of Multilayered HyperShielded Meander Line,” IEEE International Symposium on Electromagnetic Compatibility, EMC 2006, Vol. 3, Portland, OR, August 14–18, 2006, pp. 760–763.
Simbeor Electromagnetic Simulation Environment, Simberian Corporation, Seattle, WA.
Wang., T., et al., “Quasi-Static Analysis of a Microstrip Via Through a Hole in a Ground
Plane,” IEEE Transactions on Microwave Theory and Techniques, Vol. 36, No. 6, June
1988, pp. 1008–1013.
LaMeres, B., and T. S. Kalkur, “Time Domain Analysis of a Printed Circuit Board Via,”
Microwave Journal, November 1, 2000.
Kok, P., and D. D. Zutter, “Capacitance If a Circular Symmetric Model of a Via Hole Including Finite Ground Plane Thickness,” IEEE Transactions on Microwave Theory and
Techniques, Vol. 39, No. 7, July 1991, pp. 1229–1234.
Johnson, H. H., and M. Graham, High-Speed Digital Design, Upper Saddle River, NJ:
Prentice-Hall, 1993.
Buck, T. J., “PWB Design and Manufacturing Considerations for High Speed Digital Interconnection, “Applications Material, Dynamic Details Incorporated, undated.
14.6
Main Points
[24]
[25]
[26]
[27]
[28]
[29]
[30]
197
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Engelmaier, W., “Non-Functional Lands: Keep Them Vs. Remove Them,” Global SMT &
Packaging, June/July 2006, pp. 48–50.
Birch, B., “Discussion on Non Functional Pad Removal/Backdrilling and PCB Reliability,”
PWB Interconnect Solutions Inc., Applications note, undated.
Neu, T., “Designing Controlled Impedance Vias,” EDN, October 2, 2003, pp. 67–72.
Simbeor Application Note #2009_05, “Designing Localizable Minimal-Reflection ViaHoles for Multi-Gigabit Interconnects,” Simberian Corporation, Seattle, WA, April, 2009.
Simbeor Application Note #2007_08, “Design of Optimal Differential Via-Holes For
6-Plane Board,” Simberian Corporation, Seattle, WA, October 2007.
Wang, Y., and W. Hong, “Design of Differential Vias in High Speed Digital Circuit,” Microwave Conference Proceedings, APMC 2005 Asia-Pacific Conference Proceedings, Vol.
5, December 4–7, 2005, p. 4.
C H A P T E R 15
Identifying Common Signal Integrity
Problems
15.1
Introduction
Previous chapters have covered the circuit and field theory necessary for understanding the ideas fundamental to signal integrity. This chapter and Chapters 16
and 17 focus on identifying, solving, and calculating results for many common
signal integrity problems.
A select group of questions that a project manager or group leader should
consider while reviewing a signal integrity analysis is presented in this chapter.
The material can also be used when developing interview questions for signal integrity engineer candidates. The previous chapters can be referred to for detailed
explanations.
The novice signal integrity engineers will find this chapter helpful as they review their own work or as they help prepare a project plan.
This chapter cannot cover all situations or all possible solutions. Instead, it is
a tool to help think about the kinds of problems and hidden assumptions that can
become buried in a complex signal integrity task. The term ASIC is used here to
mean any integrated circuit that is connected to a microstrip or a stripline trace.
15.2 Questions to Ask When Reviewing PCB Stackup
•
Is the board thickness impacting the via size?
The minimum allowable via diameter is determined by the overall board
thickness and the ability of individual fabrication shops.
•
Vias that are too large cannot be placed in the pin field of small geometry
packages (such as dense BGA pin fields). This can require the use of higherend processing such as blind, buried, or micro-vias.
Does the design require thin dielectrics?
•
•
•
If thin core dielectrics are used to increase the capacitance between power
and ground planes, has a careful decoupling capacitor study been per-
199
200
Identifying Common Signal Integrity Problems
•
formed showing this technique is cost effective compared to discrete
capacitors?
•
Are the cores so thin that the number of shops capable of fabricating the
board is limited and thus jeopardizing second source opportunities?
Is an advanced laminate system being specified?
•
•
•
Does the stackup include many power planes?
•
•
•
•
This can be an effective way of adding routing layers with only slight increase to the stackup thickness.
Coupling between layers is minimized when signals on alternate layers are
run at right angles to one another (for instance, north-south on one layer
and east-west on the adjacent layer).
Careful signal integrity analysis is required to determine the amount of
crosstalk a bus will experience.
Are the proper voltage planes used to form stripline and microstrip?
•
•
Often routing density and stackup improvements can be had by routing
the majority of critical (often impedance controlled) signals on specially
designed layers. This also allows the signal return paths to be carefully
designed.
Are dual stripline layers used to maximum benefit?
•
•
If so, can power planes be removed from the stackup (reducing cost and
overall thickness) and can components be rearranged to allow one or more
planes to be split and shared between more than one voltage?
Are the most critical signals segregated to a specific signaling layer (or layers)?
•
•
If so, do the signal integrity simulations show an improvement in overall
board thickness, signal loss, crosstalk, or signal routing density significant
enough to warrant the added cost? If not, why isn’t FR4 (or an enhanced
FR4 system) used?
Can the chosen laminate be used by multiple fabrication shops to manufacture the circuit board? If not, do the electrical benefits outweigh the risk
of limited second sourcing?
A voltage plane unrelated to the signaling voltage must be carefully analyzed and modeled. Otherwise, the actual circuit will experience noise not
predicted by simulation.
Are mounting holes located in a critical routing area?
If so, verify that signaling has not been impaired by routing too close to the
hole or to a metallic screw.
Are sharp rise time or critical signals routed close the board edge?
•
•
•
•
If so, these signals may experience increased jitter and poor impedance
control, and the board may have increased EMI.
What is the via/antipad strategy?
•
Has the layout designer been allowed to select the via and antipad sizes,
or have these been designed? Often the layout designer selects the via
size based on layout and routing considerations. These may be optimum
15.3 Questions to Ask When Reviewing Crosstalk Simulations
•
•
•
•
Has an analysis been done trading off ASIC driver strength and termination strategies and trace impedance?
Lower-impedance traces result in a thinner overall stackup and make the
signals more immune to crosstalk, but increase driver current.
Higher-impedance traces require less drive current and are more immune
to conductor loss but more sensitive to dielectric losses.
How was the trace width arrived at for each signal or signal class?
•
•
•
•
•
•
from a manufacturing standpoint, but may not provide the best electrical
behavior.
Large antipads can remove extensive metal from the power or ground
plane, especially in dense pin fields such as under BGAs. This can result in
routing and (with high-power devices) power distribution problems.
How was the trace impedance value selected for each layer?
•
•
201
Often the minimum trace width allowed by the manufacturer is chosen
first, before any signal analysis is performed, but in some circumstances
a wider width is preferable. For instance, widening the trace reduces conductor loss, and the crossover frequency where the dielectric loss becomes
larger than the conductor loss will change.
Wider traces are easier to manufacture and generally result in less scrap
(lowering costs).
Wide traces have lower inductance and larger capacitance than narrow ones. This is advantageous when routing power to terminators, for
example.
A trace that is too narrow will limit the number of second sources capable
of manufacturing the board in volume.
Narrow traces often have more variability in shape (and therefore impedance and resistance) than do wide traces.
How was the minimum trace spacing determined?
•
•
Have crosstalk simulations (see Section 15.3) been performed that demonstrate proper operation with the selected spacing?
Spacing that is too small will limit the number of second sources capable
of manufacturing the board in volume.
15.3 Questions to Ask When Reviewing Crosstalk Simulations
•
Does the I/O model produce the fastest signal rise time?
•
•
The fast (sometimes called “best”) process corner produces the fastest rise
time.
CMOS drivers produce the fastest rise time when the lowest permitted
temperature and highest permitted voltage are used with the fast transistor models.
•
Does the circuit board trace model include both conductor and dielectric
losses?
•
Has the termination resistance and voltage been set to the worst case?
202
Identifying Common Signal Integrity Problems
•
•
Have the effects of more than one neighboring trace been included in the
simulation?
•
•
•
•
To avoid long-term or catastrophic damage to the I/O circuits, the input
voltage must not exceed the maximum allowed.
Is the induced voltage large enough to cause the receiver output to change
state when the effects of power and ground noise caused by a nearby driver
switching are present?
•
•
Unrealistically high voltages can be reported if these are not included.
The effects of multiple reflections will not be properly reported if the
clamps are missing.
Does the induced voltage exceed the safe values allowed for the process
technology?
•
•
Not including input capacitance affects the receiver response time and the
way in which reflections are formed, which will change the crosstalk magnitude and timing.
Not including output capacitance affects the driver rise time.
Do the receiver models include the effects of clamping circuits?
•
•
By mistakenly focusing on far-end crosstalk, problems at the near end can
be overlooked.
Do the receiver circuit models properly include parasitic capacitances on the
inputs and outputs?
•
•
If the timing is just so, a signal from a bus unrelated to the bus being simulated can add to the total crosstalk if they are routed close to one another.
Is near-end or far-end crosstalk the bigger concern in this design?
•
•
Unless the neighbors are widely spaced, the influence of multiple neighbors
is required to obtain the worst-case crosstalk.
Was it necessary to include the contributions of signals from other buses?
•
•
Using nominal values will produce the least crosstalk.
If so, is the system timing such that this can be ignored?
Is it possible for the bus to be put in a high-impedance (nondriven) state?
•
•
If unterminated, this case will probably yield the highest crosstalk.
Do the induced voltages remain within the safe range for the process
technology?
15.4 Questions to Ask When Reviewing Signal Quality Simulations
The task here is to insure that signal quality has been tested under all possible valid
worst-case conditions. Poor signal quality may be caused by a signal that is driven
too weakly (underdriven) or by one that is driven too hard (overdriven). The signal
integrity simulations must test for both cases.
Underdriven signals only gradually reach the receiver switch point and may not
provide adequate noise margin. In high-performance systems underdriven signals
are usually an unintentional consequence of the design and are seen as errors to
15.4 Questions to Ask When Reviewing Signal Quality Simulations
203
be corrected. However, sometimes signals are intentionally underdriven to remove
high-frequency harmonics as a way to reduce EMI and to save power in low-power
applications.
Overdriven signals are sharp edged and overshoot past the static high value.
The overshooting may be severe enough for the signal to momentarily fall (to ring
back) below the receiver switch point before it settles at the steady state value. For
large overdrive conditions the signal may oscillate multiple times between successively smaller overshoot and ringback values before it finally comes to rest at a
steady state value.
Therefore, the following questions should be asked when reviewing signal integrity simulations:
•
What analysis was performed to guarantee that both the worst-case underdrive and worst-case overdrive situations have been identified?
•
•
•
•
In some systems the difference in signal quality between these two will be
small, but usually the difference is demonstrable.
In CMOS systems underdrive is usually the worst when the slowest process corner is used at the highest temperature and lowest power supply
voltage and the highest series resistance (if the signal is series terminated).
In CMOS systems worst-case overdrive usually occurs when the fastest
process corner is used at the highest voltage and lowest temperature. If
series terminated, the minimum resistance value is used.
If the signal experiences ringback, has its impact on system timing been thoroughly analyzed?
•
For example, a worst case may be for a data strobe or clock to overshoot
(and reach the switch point sooner than expected) while the data signal
its clocking is underdriven and so only gradually reaches the switch point.
•
If the signal overshoots, does the peak value remain within the manufacturer’s safe values specification?
•
If the signal is underdriven, how has the effect of power supply noise or noise
on adjacent signals been tested?
•
•
Have all the sources of signal coupling been identified and included in the
simulation?
•
•
Since underdriven signals only gradually reach the receiver switch point,
they are especially vulnerable to false triggering caused by crosstalk or by
noise coupled to the signal from the power supply.
See the earlier discussion on crosstalk.
Have the worst-case data patterns been identified?
•
The worst-case switching pattern depends on the exact circumstances of
the design, but in general, a simple 1/0 “square wave” pattern is usually
not the worst case. Often the worst pattern involves timing the transmission of a bit so that it combines with energy still present on the line, but
the “lone one” [1, 2] can also be significant, especially in situations where
the line is lossy.
204
Identifying Common Signal Integrity Problems
•
•
•
Bidirectional buses are especially challenging to simulate and the interaction when different drivers (presumably connected with unequal lengths of
trace) are asserting the bus must be tested under all combinations of timing
and impedances variations. Monte Carlo [3, 4] or some other statistical
method is valuable in reducing the number of simulations, but some SI
CAD can be configured to run all possible permutations. Software (either
acquired as part of the SI CAD suite or a simple script written by the SI engineer) is critical in sifting through the data and selecting the worst cases.
Often bidirectional buses have multiple worst-case patterns that depend
on the direction of data flow and trace lengths. The SI engineer must be
persistent to find all the worst-case patterns and not to stop looking once
one or two have been identified.
Has SSN been properly accounted for?
•
A common mistake when checking the switching behavior of an ASIC is
to assume that the worst-case “signal noise” occurs when all the I/O pins
drive simultaneously, often in the same direction. While this usually creates
the most crosstalk and noise coupling, it does not necessarily create the
worst reflections on the circuit board. In fact, the current demand when
many drivers simultaneously switch may cause the power supply inside the
ASIC feeding the I/O drivers to momentarily drop. In CMOS I/O drivers
this current starvation retards the driven rise time. The net effect is that the
rise time may be fastest when only one (or a few) drivers switch and may
be the slowest when many simultaneously switch.
15.5 Questions to Ask When Reviewing Prelayout Simulations
•
What tests have been done to validate each component model?
•
•
•
•
What has been done to demonstrate that models from various vendors will
function in the same simulation?
•
•
Circuit-level models sometimes require unique settings and parameters to
run properly. The simulation will not operate correctly if the same parameters must be set differently for various models coexisting in the simulation.
What is the strategy for determining the worst-case simulation conditions?
•
•
•
A common mistake is to accept the vendor’s circuit model for each component without independently testing it for compliance.
At a minimum the model should be run in a simple test circuit to verify
that it is operating as expected.
In critical designs the model should be validated with actual laboratory
measurements before beginning detailed signal integrity simulations.
Realistic estimates of the junction temperatures must be made. If the die
temperature of interconnected ASICs is not the same, a strategy addressing
the behavioral difference these causes must be developed.
Have the effects of power supply noise been accounted for?
Have trace lengths been estimated properly, and what assumptions have been
made regarding component placement?
15.6 Questions to Ask When Reviewing Postlayout Simulations
•
•
•
Component models that include inductance are necessary to obtain the
correct high-frequency behavior.
Models of vias and any trace connecting the terminators to a power or
ground system must also be included.
Have the worst-case patterns been identified?
•
•
•
•
Proper estimates of component placement are critical for accurate simulation results because the amplitude and timing of reflections depend on the
trace lengths.
What assumptions have been made regarding termination placement and the
inductance connecting them?
•
•
205
Were a series of manual simulations run, or was a statistical approach
(such as Monte Carlo) used? What has been done to demonstrate that the
worst-case patterns have actually been found?
Do the simulations include signal loss, or have lossless lines been assumed?
Do the simulations include crosstalk effects, or have their effects been accounted for by including a noise “adder”?
Have models been created for the vias?
•
•
If so, what stackup assumptions have been made?
What process is in place to ensure that the vias in the actual product match
the mechanical details used in the electrical model?
15.6 Questions to Ask When Reviewing Postlayout Simulations
•
Has the layout artwork been fully checked for compliance with the layout
rules?
•
Have timing and signal quality simulations been performed with data from
the fully routed circuit board?
•
•
Is power supply decoupling adequate and is the capacitor placement and
layout optimal?
•
•
If long enough, stubs will create reflections that can interfere with proper
signaling.
Do the postroute SI simulations include the actual number of layer changes?
•
•
Verify that ASIC manufacturers’ guidelines have been followed for the
placement and values of decoupling capacitors.
Are previously unsimulated stubs present on any net, especially high-speed
nets?
•
•
Has the timing on all critical nets been verified?
Vias increase signal loss and can cause reflections. Ensure that the actual
layout matches the simulation model.
Is the location of series or parallel termination resistors acceptable?
•
•
Does it match the simulations?
Do the simulations include the range of trace lengths and the impedance of
the traces connecting the terminators?
206
Identifying Common Signal Integrity Problems
•
If a fly-by trace is used to connect a parallel terminator to a load, has that net
been analyzed for coupling and crosstalk?
•
Are any vias shared?
•
Vias connecting decoupling capacitors should never be shared: each decoupling capacitor should be connected to power and ground through
individual vias.
15.7 Questions to Ask When Reviewing a Termination Strategy
•
Has termination located on the ASIC die (ODT) been investigated?
•
•
•
If a parallel termination strategy is used, in this application is it more costeffective to use a resistor pair, or should a Thevenin resistor/supply be used?
•
•
•
This is generally superior to termination placed on the circuit board.
It may not allow as much flexibility in make adjustments.
The two-resistor scheme is simple and does not require creating and routing a separate supply voltage.
Often the single resistor Thevenin termination can be placed closer to the
load and is especially attractive in large, densely packed designs.
Does the simulation model account for the proper location of the series termination resistors?
•
The model should include the actual distance from the I/O pin to the termination resistor, and the impedance of that trace (which is might be higher
than the remainder of the signal trace) should be included.
15.8 Questions to Ask When Reviewing PCB Layout Design Rules
•
Have the recommendations of the ASIC manufacturer been incorporated
into the design rules?
•
•
The quality and practicality of ASIC vendor-supplied signal and power
integrity rules vary greatly. Nevertheless, the ASIC manufacturer’s goals
should be observed when developing signal integrity rules for an end
product.
Has an ASIC manufacturer’s reference design been studied?
•
Valuable insight regarding the unique layout requirements of an ASIC can
be obtained by reviewing the reference design supplied by the manufacturer. Especially valuable are stackup and decoupling (or other power supply filtering) information, but termination and trace impedance factors can
also be discovered.
15.9 Questions to Ask When Reviewing ASIC Driver Selection Choices
•
Have drivers with various drive strengths and rise times been tested?
15.9 Questions to Ask When Reviewing ASIC Driver Selection Choices
•
•
•
•
Have the circuit board traces’ odd/even modes been properly simulated?
Does the simulation include parasitic inductances, capacitance, and resistance for the ASIC package and socket?
Does the simulation include some portion of the power distribution network on the ASIC and on the circuit board?
Has the case of only one driver switching been simulated?
•
•
Has the weakest driver with the least aggressive rise time that still meets
the timing and receiver noise budgets been selected?
Has the effect of many drivers simultaneously switching been simulated?
•
•
207
Does the signal rise time become so fast that transmission line effects are
unacceptable?
Have the simulations been performed across the full range of voltage, process, and junction temperatures?
•
Does the simulation include the effects of manufacturing tolerance and
temperature changes on the board level termination resistors and termination supply voltages?
References
[1]
[2]
[3]
[4]
Dally, W. J., and J. W. Poulton, Digital Systems Engineering, Cambridge, U.K.: Cambridge
University Press, 1998.
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Hall, S. H., and H. L. Heck, Advanced Signal Integrity for High-Speed Digital Designs,
New York: John Wiley & Sons, 2009.
Li, P. M., Jitter, Noise, and Signal Integrity at High Speed, Upper Saddle River, NJ: PrenticeHall, 2007.
CH A P T E R 16
Solving Common Signal Integrity
Problems
16.1
Introduction
This chapter outlines strategies for resolving several common signal integrity
problems.
It is best to solve these problems during the simulation analysis phase, before
the circuit board is manufactured. Often signal integrity errors are systematic in
nature. All signals in a net class are usually misrouted in the same way and so they
all have the same weaknesses. However, it is not unusual for the “noise problem”
to initially appear to be on only a few signals. Once the problem on these signals is
corrected, the weaknesses in the remaining signals become apparent, and they too
must be fixed.
The general rule is that once a layout problem has been identified for a specific
signal, that fix must be applied to all of the signals in that same net class. If the
problem has been found on actual hardware (rather than in simulation, before the
board has been manufactured), this often means that the circuit board must be
relayed out.
The suggestions listed here are intended to identify the problem’s underlying
cause and are not an exhaustive list of remedies. They can be used when performing signal integrity simulations and during laboratory debug. They are also useful
for discussion points when interviewing signal integrity candidates.
This chapter broadly summarizes the material presented in earlier chapters.
16.2
Reducing Crosstalk
Crosstalk can be minimized by reducing the parasitic coupling between traces or
by reducing the signal strength of the aggressors. Often a combination of these is
used in practice.
In particular, to reduce crosstalk:
209
210
16.3
Solving Common Signal Integrity Problems
•
Increase the spacing between traces.
•
Move critical signals to a separate routing layer.
•
Ensure that the signal reference planes also connect to the I/O drivers (do not
use an unrelated voltage plane as a reference).
•
Add a properly grounded guard trace.
•
Minimize the coupled length.
•
Make the victim trace as low impedance as is practical.
•
Lengthen the signal rise time (make it slower).
•
Reduce the aggressor signal voltage swing.
•
Properly terminate the aggressor and victim traces.
•
Adjust timing so that the crosstalk event does not cause false triggering (this
reduces the effect but does not eliminate the actual coupled voltage).
•
Adjust timing so that multiple aggressors do not simultaneously couple onto
a victim (this reduces the peak voltage coupled onto the victim).
Reducing Reflections
Reflections occur when the load impedance does not match the transmission line
impedance.
Termination is required if the line is electrically long. Do the following when
excessive ringing, overshooting, and ringback are present on a net that has been
terminated:
•
Verify that the manufactured trace impedance is the expected value and that
it matches the termination.
•
Perform an odd/even mode analysis to determine if the difference between
the odd/even impedances is high. If so, consider increasing the spacing between traces or adding a guard trace to control the impedance under switching conditions.
•
On multidrop nets perform a simulation analysis to determine the best way
to connect the loads and to find the optimum termination type and value.
•
Reduce the I/O cell drive strength to decrease the launched rise time.
•
Shorten the trace length.
•
If properly matching the load impedance to the transmission line impedance
is not possible (for example, a large lumped capacitive load), eliminate multiple reflections by ensuring that the driver impedance closely matches the
line impedance.
•
Make termination more effective by breaking up a multiload net into several
identical copies, each of which is driven from separate pins on the ASIC.
•
Ensure that series termination resistors are placed close to the driving pin.
•
Ensure that parallel terminations are properly connected to the load.
16.4 Reducing and Eliminating Simultaneous Switching Noise (SSN)
211
•
Ensure that the signal reference planes also connect to the I/O drivers (do not
use an unrelated voltage plane as a reference).
•
Ensure that adequate decoupling is placed between power and ground planes.
•
If a Thevenin termination is used, ensure that the Thevenin voltage is stable.
It must not oscillate and should have little droop even when many signals
connected to it simultaneously switch.
16.4 Reducing and Eliminating Simultaneous Switching Noise (SSN)
Eliminating SSN from a poorly designed circuit board can be difficult and, in some
cases, impossible. It is best to analyze SSN effects before prototype hardware is fabricated, especially that of the micropackage and any connectors (including sockets).
The following circuit board fixes can be helpful when an integrated circuit
experiences high SSN in those cases where little or no SSN analysis has been done.
This list is also useful during debug to help identify SSN.
•
Use power and ground planes to connect the ASIC to the power supply.
•
Provide adequate decoupling capacitance adjacent to the IC pins. This is
extremely critical.
•
Reduce the number of pins that simultaneously switch.
•
Arrange for some signals to always switch in the opposite direction (oddmode parity, for instance) or stagger the timing so that some signals switch
later than others.
•
Use differential signaling.
•
Use a less powerful (lower current drive) I/O cell.
•
In CMOS systems decrease the power supply voltage and increase the junction temperature. This will reduce SSN but will also adversely affect timing.
•
Increase the transmission line impedance, since this reduces the launched
current.
•
Use series termination on the circuit board since this reduces the launched
current.
16.5 Improving Inadequate Timing Margins
Assuming that the root cause for the poor timing margin is signal integrity related
and that it does not involve poor logic design, one or more of the following circuit
board fixes can be helpful:
•
Guarantee that ringback and reflections are not eroding setup or hold times.
•
Use a higher current drive I/O cell, remove series termination, and provide
parallel termination at the load (but these may increase crosstalk and simultaneous switching noise).
212
Solving Common Signal Integrity Problems
•
If receiver setup time is inadequate, reduce the signal delay by increasing the
driver I/O power supply voltage and decreasing the driving ASIC’s temperature. Shorten the trace length and use microstrip rather than stripline.
•
If receiver hold time is inadequate, increase the signal delay by lowering the
driver I/O power supply voltage and/or increasing the driving ASIC’s temperature. Lengthen the trace.
•
Use serpentine traces to adjust the timing of a strobe relative to its data.
•
Small adjustments in timing can be made by switching to a laminate system
with a different dielectric constant (but this will change the trace impedance
unless the stackup is also adjusted).
16.6 Correcting Intersymbol Interference (ISI)
ISI is caused by a combination of dispersion, crosstalk, signal reflections, and jitter.
Ways to reduce crosstalk and reflections are covered earlier in this chapter. In addition to those general guidelines, some specific steps to take include:
•
Reduce the frequency content of the signal by degrading the signal rise time
and slowing the data rate to increase the bit time.
•
Decrease microstrip and stripline conductor losses by using wide, high-impedance traces (but this will increase the stackup height and via length).
•
Reduce dielectric losses in microstrip (but not stripline) by using wide, highimpedance traces (but this will increase the stackup height and via length).
•
Use a low loss tangent laminate system (but this may not be cost-effective for
price-sensitive products).
•
Use a low dielectric constant laminate system.
•
Adjust I/O driver precompensation and impedance control circuits. These
circuits automatically correct for frequency-dependent losses and they better
match the I/O cell impedance to the line impedance.
•
When working with cables or very long traces, consider adding passive
equalization [1–5].
•
Minimize the number of vias in the signal chain since vias can create reflections and can act as lowpass filters that attenuate high-frequency harmonics.
•
Use impedance matched vias.
•
Ensure that the differential impedance is within specification and that the
proper termination is used.
•
Ensure that nets where ISI is of concern are routed well away from other
traces, away from cutouts in the planes, are not near the board edge, and are
away from unrelated vias.
16.7 Steps to Take When Circuit Board Traces Are Too Lossy
213
16.7 Steps to Take When Circuit Board Traces Are Too Lossy
At high frequency the total interconnect losses are improved by reducing dielectric
losses and conductor losses (losses caused by the resistance of the traces).
•
Reduce the trace length, since loss increases exponentially with length.
•
Use a 2D field solver to determine if the major contributor to the total loss is
due to conductor loss or dielectric loss. Do not use a high-cost low dielectric
loss laminate system if the loss is predominately due to the conductor.
•
If conductor losses dominate, consider using low surface roughness copper.
•
If dielectric losses dominate, consider using a low loss tangent laminate system (but unless very high-performance laminate systems are used, this only
slightly reduces dielectric loss and so may not be acceptable in cost-sensitive
applications. However, this may be the only viable solution when signaling
at high speeds).
•
Increase the trace width, but note that thicker traces will not improve highfrequency losses.
•
Use high-impedance traces. This is especially effective in lowering dielectric
and conductor losses in wide microstrips, but conductor losses in striplines
will also improve.
•
Consider routing critical signals as microstrip. This can be especially effective
when signaling at high-data rates since the dielectric loss usually dominates at
high-frequency, and dielectric losses are smaller for microstrip than stripline.
However, because manufacturing tolerances are generally looser on the outer
layers, impedance control may not be as good and reflections may increase.
•
If the traces are microstrip, consider removing the solder mask and treating
the exposed traces with a surface finish instead. This creates an exposed
microstrip rather than an embedded microstrip and reduces dielectric loss.
However, this can cause assembly problems. Be sure to check with manufacturing to verify that this option is acceptable.
•
If only a few critical signals are experiencing unacceptable loss, consider
using a mixed laminate system where a high-performance laminate is used
on one layer in the stackup, while the remaining layers are traditional FR4
material. By segregating the critical stripline routes to the high-performance
layer, the loss is reduced for those signals and the remaining lower-cost layers
are used to route the majority of the signals.
•
Verify that the signals are referenced to the proper power plane.
16.8 Options to Reduce Circuit Board Thickness
Using more routing layers than necessary is the most typical cause of thicker than
desired circuit boards. To reduce board thickness:
214
Solving Common Signal Integrity Problems
•
Use dual stripline (but this can cause crosstalk unless signals are routed
orthogonally).
•
Use thin core laminate technology between the power and ground planes (but
this may not be cost-effective in price-sensitive products).
•
Improve routing density to reduce the number of routing layers by using narrow traces (but this increases signal loss on long traces and may reduce the
number of vendors capable of second sourcing the board).
•
Use low impedance traces on inner layers (but this increases the current
driven by I/O devices and may create SSN problems).
•
Use a low dielectric constant laminate system (this works best when the
board has many layers).
•
Consider using half-ounce or quarter-ounce copper for traces and planes (this
only provides a slight reduction; the loss and current handling effects must
be analyzed).
•
Consider removing some power planes by splitting them for use by more
than one supply voltage. This may require components to be rearranged on
the board.
•
Consider replacing power planes with power islands located on the board
surface.
•
Consider routing slow, noncritical stripline signals on a layer where one of
the planes is not the I/O voltage. For acceptable operation this requires careful design and analysis.
•
If the board is too thick to accommodate a particular connector at the edge
of a board, the circuit board fabricator can reduce the thickness in selective
regions. This technique is not always cost-effective, but it allows the stackup
to accommodate many planes and routing layers while simultaneously being
thinner at the board edge.
16.9 Steps to Take When There Are Not Enough Routing Layers
•
Use dual stripline to increase the number of routing layers without adding
too much thickness.
•
Improve routing density to reduce the number of routing layers by using narrow traces (but this increases signal loss on long traces and may reduce the
number of vendors capable of second sourcing the board).
•
Use low-impedance traces on inner layers and decrease the separation between traces (but this increases crosstalk and the current driven by I/O devices and may require that the nets be terminated).
•
Consider removing some power planes by splitting them for use by more
than one supply voltage, or using power islands, and using the free layer for
routing.
16.10 Steps to Take When a Circuit Board Must Be Cost Reduced
215
•
Consider using serial signaling technology to eliminate wide, synchronous
buses.
•
Consider using a cable or flexible circuit board assembly to jumper signals
from one end of a large circuit board to another. This can remove the need to
route critical, highly sensitive signals on the board, or it can be used to route
many low-performance signals.
16.10 Steps to Take When a Circuit Board Must Be Cost Reduced
•
Send the board stackup out to several vendors to determine if it can be
simplified.
•
Determine if a slight reduction in the board dimensions will significantly
reduce cost. Laminate panels come in standard sizes and cost will be lower
if the board is small enough to allow several copies to fit within one of the
standard sizes with little waste. A slightly smaller board that allows several
copies to be placed on a larger panel will cost less than a larger board that
can only fit one copy per panel.
•
Understand what characteristics are increasing manufacturing costs of the
raw board. For instance, the need for advanced via technology or very fine
lines can drive up cost.
•
Study the decoupling very carefully. Often decoupling is not rigorously analyzed during design, and more capacitors than necessary are used. Use simulation to determine the worst-case conditions for power supply noise and
replicate that on prototype hardware. Then selectively remove decoupling
on the prototype to determine the minimum necessary for proper operation.
•
Verify that all of the terminated nets actually require termination. Removing
unnecessary termination components frees up board space (possibly allowing
the board to be smaller or to use fewer layers) and, when using parallel termination, power can be reduced, and the costs related to parts and assembly
are also reduced.
•
If controlled impedance traces are used, verify that they are necessary. Requiring controlled impedance traces increase the bare board manufacturing
cost.
•
Verify that trace impedance specifications are not overly restrictive and are
easy to manufacture. For example, many manufacturers are accustomed to
producing 50Ω single ended and 100Ω differential pairs. Specifying different values than this will usually increase the number of boards scrapped and
increase cost.
•
Verify that spacing between traces is not wider than actually necessary to
reduce crosstalk. Spacing that is too great reduces the number of available
wiring channels and may increase the number of required routing layers.
•
Investigate the use of dual stripline. This can reduce the layer count and
board thickness, which can have a cascading effect on reducing board cost.
216
Solving Common Signal Integrity Problems
16.11 Steps to Take When Data-Dependent Errors Are Causing System
Failures
•
Rule out crosstalk (refer to Section 16.2).
•
Rule out SSN (refer to Section 16.4).
•
Verify by laboratory measurement or simulation that pulse smearing caused
by dispersion (especially on long traces) is not preferentially distorting some
bits. Dispersion can occur on parallel buses and is not only an artifact of serial signaling.
•
Verify that the signal is properly terminated. Improper termination can result
in extreme voltages being present at receiver inputs, which can require successive bits excessive time to assert to valid logic levels. Additionally, signals
transitioning well below ground or well above the power rail can require a
long recovery time from the receiver before proper timing is restored.
•
Verify that data-dependent signal loss is not causing pulse distortion. If the
signal return path is inadequate, the apparent AC resistance of a trace will
increase (causing the signal loss to increase) as the return current from many
signals commingle. Review the circuit board artwork to see if many signals
are returned by a narrow strip of ground plane, or if multiple signals are
forced into a small, isolated channel. If a power plane is one of the return
paths, verify that it is the I/O voltage and is not unrelated.
•
Verify that connectors have adequate numbers of ground pins (ideally one
ground per signal, but fewer in practice with the actual number required being determined by simulation or laboratory measurement). If the connectors
are used to signal between boards, verify that the connector ground pins
are connected by low-impedance paths to the signal return planes on their
respective boards.
References
[1]
[2]
[3]
[4]
[5]
Couch, L. W., Digital and Analog Communication Systems, 5th ed., Upper Saddle River,
NJ: Prentice-Hall, 1997
Bennett, W. R., and J. R. Davey, Data Transmission, New York: McGraw-Hill, 1965.
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood, MA: Artech House,
2004.
Proakis, J. G., Digital Communications, New York: McGraw-Hill, 2000.
Federal Telephone and Radio Corp, Reference Data for Radio Engineers, 3rd ed., New
York: Knickerbocker Printing Corp., 1949.
C H A P T E R 17
Calculating Trace and Plane Electrical
Values
17.1
Introduction
With the formulas presented in this chapter, the signal integrity engineer can estimate the impedance, time of flight, loss, and circuit values (R, L, C) for traces and
the circuit values for planes. Ways to calculate conductance (G) appear in Chapter
7 and are not repeated here.
These are very helpful in estimating the parasitic values of traces without resorting to a field solver. In fact, the results from these equations can be used as a
first-order check of the results from field solvers when validating circuit models.
Some of the equations (especially for impedance and loss) are too complicated
to be performed reliably by hand on a calculator. However, they are easily solved
with numerical software such as MATLAB [1] or MathCAD [2]. Spreadsheets can
also be used, or the equations can be coded in a simple programming language
(such as C or even a scripting language such as PERL, particularly for the simpler
equations).
The dimensions for stripline and microstrip traces used in the formulas are
shown in Figure 17.1.
17.2 How to Convert from Decibel Loss to Signal Swing
Chapter 8 showed how to calculate loss in decibels (dB). That equation is repeated
here, in (17.1):
⎛ V
⎞
Gain in dB = 20 × log10 ⎜ received ⎟
⎝ Vtransmitted ⎠
(17.1)
For example, the gain is −6 dB when Vreceived is 0.5V and Vtransmitted is 1V. Since
the sign is (−), this value represents a loss.
217
218
Calculating Trace and Plane Electrical Values
tsm
t
w
b
εr
h
Stripline
(a)
εr
Microstrip
(b)
Figure 17.1 (a) Stripline trace is centered between two return planes, separated by distance b.
(b) Microstrip is raised h above the return plane and the solder mask thickness is tsm. The traces have
a width of w and a thickness of t, and the dielectric constant is εr.
Some signaling specifications give the total loss budget for the link in decibels.
In those situations converting from loss in decibels to a voltage ratio is useful for
understanding the required launched and received voltages.
Solving (17.1) for the ratio of Vreceived to Vtransmitted produces:
dB
⎛ Vreceived ⎞
20
=
10
⎜⎝ V
⎟
transmitted ⎠
(17.2)
For example, the Vreceived to Vtransmitted ratio is 0.71 when the loss is 3 dB. Here
again, a (−) sign is used to show loss. If the minimum valid receiver voltage is 0.8V,
with a 3-dB transmission line loss, the transmitter must launch 1.1V for the link to
properly operate.
17.2.1 Nepers
Rather than using common logarithms to express loss in decibels, natural logarithms can be used to express the loss in nepers (Np) [3, 4]. Nepers are not as
frequent as decibels in signal integrity work, but the conversion shown in (17.3) is
useful when it is necessary to convert between the two.
dB = Np × 6.868
(17.3)
For instance, a loss of 27.5 dB can also be represented as a 4-Np loss.
17.3 Estimating DC Resistance
Equation (17.4) shows how the DC resistance of a trace or plane depends on the
metals resistivity (ρ, see Chapter 5), cross-sectional area (A), and length.
Rdc =
ρ
× Length
A
The trace dimensions are shown in Figure 17.2.
(17.4)
17.4 Finding Inductance and Capacitance When the Physical Dimensions Are Not Known
219
Length
w2
t
w1
Figure 17.2 Trace dimensions for calculating DC resistance. Set w1 = w2 for rectangular traces.
The resistance of rectangular or trapezoidal copper traces in ohms per inch or
per meter can be found by using:
Rdc =
Krdc
t ( w1 + w2 )
(17.5)
Set Krdc to 1.41 to find the resistance in ohms per inch length when t, w1, and
w2 are in mils. To find the resistance per meter length, use 352 for Krdc and set t,
w1, and w2 to micrometers.
For a perfectly rectangular trace, make w1 equal to w2. For example, to determine the resistance of a 5-inch-long perfectly rectangular 5-mil-wide, 0.65-milthick trace, Krdc is 1.41, t is 0.65, and w1 and w2 are each 5. The equation yields
0.216 per inch, making the 5-inch-long trace resistance 1.1Ω.
17.4 Finding Inductance and Capacitance When the Physical
Dimensions Are Not Known
The impedance (Zo) and delay (tpd) equations (presented in Chapter 6) can be
simultaneously solved to find the trace inductance (17.6) and capacitance (17.7).
L = Zo × tpd
C=
tpd
Zo
(17.6)
(17.7)
220
Calculating Trace and Plane Electrical Values
For example, the inductance L of an 86Ω transmission line having a delay of
5.6 ns/meter is found from (17.6) to be 482 nH/meter, and from (17.7) its capacitance is 65 pF/meter.
17.5 Finding Stripline Inductance and Capacitance When Impedance
Is Known
Because the dielectric is homogeneous in stripline (see Chapter 7), the inductance
and capacitance equations can be determined for that specific case when only the
impedance is known.
The inductance of a stripline is found with (17.8). Set K1 to 0.085 to obtain the
inductance in nH/inch. Use 0.033 to obtain the inductance in nH/cm.
Lsl = Zo × K1 εr
(17.8)
The capacitance of a stripline is calculated with (17.9). Set K2 to 85 to obtain
the inductance in pF/inch. Use 33 to obtain the inductance in pF/cm.
Csl =
K 2 εr
Zo
(17.9)
For example, the capacitance of a 50Ω stripline of any width or thickness is 3.4
pF/inch length when the dielectric constant (εr) is 4.
17.6 Finding Stripline Inductance and Capacitance When Trace
Geometry Is Known
Inductance of a stripline trace can be estimated with (17.10). This equation has
been simplified and curve-fitted [3] from [5]. It assumes that the trace is centered
between the two planes. Since it does not account for trace thickness, it is only accurate to within about 20% of the actual value.
L = K3 ×
b
w+b
(17.10)
where K3 is 16 for results in nH/inch, or 6.3 for nH/cm.
The reciprocity principle [3] can be used to find the capacitance of stripline
(but not microstrip) from the inductance calculated in (17.10). Doing so yields
(17.11), and since it is derived from (17.10), it has the same level of accuracy.
⎛ w + b⎞
C = K 4 × εr × ⎜
⎝ b ⎟⎠
(17.11)
17.7 Finding Microstrip Inductance and Capacitance When Trace Geometry Is Known
221
where K4 is 0.45 for results in pF/inch or 0.18 for results in pF/cm.
Because in both of these equations the units cancel, K3 and K4 can have different units than h and w.
For example, the capacitance of a 5-mil-wide stripline centered between two
planes spaced 66 mils apart when εr = 4.0 is 0.77 pF/cm (K4 = 0.18, w = 5, h = 66).
This is 15% higher than the actual value of 0.67 pF/cm.
17.7 Finding Microstrip Inductance and Capacitance When Trace
Geometry Is Known
Often the width, thickness, and impedance of a microstrip are known, but the time
of flight is not. In that case the (17.12), (17.13) and (17.14) can be used to estimate
the inductance and capacitance.
17.7.1 Microstrip Inductance
Inductance of a microstrip trace not covered in solder mask can be estimated (without accounting for trace thickness) with (17.12). This has been simplified and
curve-fitted [3] from [5].
h⎞
⎛
L = K5 × ln ⎜ 6.3 × ⎟
⎝
w⎠
(17.12)
Set K5 to 5.1 to obtain the inductance in nH/inch or to 2 for nH/cm. Because
the units cancel, K5 can have different units than h and w.
For example, the inductance of a 5-mil-wide microstrip spaced 11.5 mils above
the ground plane is 5.35 nH/cm (K5 = 2, w = 5, h = 11.5). This is 7% lower than
the actual value of 5.46 nH/cm.
17.7.2 Microstrip Capacitance
Because the dielectrics are not homogeneous, it is not proper to use the reciprocity
principle to find the capacitance when the inductance is known, as was done with
stripline.
Equation (17.13) is simplified from [6] and used to find the capacitance once
the effective dielectric constant has been found with (17.14) [7] (repeated here from
Chapter 7).
C=
εr _ eff
K6 × εr _ eff
⎛ 8h w ⎞
ln ⎜ + ⎟
⎝ w 4h ⎠
⎛
⎞
⎟
εr + 1 ⎜ εr − 1
1
=
+⎜
×
⎟
2
10h ⎟
⎜ 2
1+
⎜⎝
⎟
w ⎠
(17.13)
(17.14)
222
Calculating Trace and Plane Electrical Values
Because the units cancel, these two equations may be used for traces dimensioned in either metric- or inch-based units. Set K6 to 1.414 to calculate the capacitance in pF/inch. Use 0.556 for K6 to determine capacitance in pF/cm.
For example, the capacitance of a 5-mil-wide microstrip-spaced 11.5 mils above
the ground plane when εr is 4.0 is 0.535 pF/cm [K6 = 0.556, w = 5, h = 11.5, and
from (17.14) εr_eff = 2.806]. This is 3% lower than the actual value of 0.55 pF/cm.
17.8 Estimating Inductance and Capacitance of a Plane
The following parallel plate capacitance and inductance equations can be used to
estimate the inductance and capacitance of power and ground planes (see Figure
17.3). Because these equations do not take fringing into account, they are most
accurate when w is extremely greater than h. This makes them unsuitable for use
with traces.
17.8.1 Inductance
Parallel plate inductance is found with:
L = K7 ×
Length
×h
w
(17.15)
w
h
Le
ng
th
(a)
w
h
Le
ng
th
(b)
Figure 17.3 Parallel plate (a) inductance and (b) capacitance present between parallel plates. The
selection of width and length matters in determining the inductance, but the capacitance is determined by the area.
17.9 Calculating Trace Loss from a Circuit Model
223
To obtain the inductance in nH/inch, set K7 to 32; set it to 12.6 for nH/cm.
The equation is written this way to show the importance of the length-to-width
ratio in determining the inductance. As with the capacitance, the h and w units cancel, so L can be found in terms of either inches or meters for h and w in any units,
provided that the units for K7 and length match.
For example, the inductance of a 10-cm-long, 50-mil-wide plane placed on a
1-mil-thick block of laminate is 2.5 nH (K7 = 12.6, Length = 10, w = 50, h = 1)
when the current is injected along the plane’s wide edge. This is 0.25 nH/cm. Because the length is in centimeters, 12.6 was chosen for K7 rather than 32.
The inductance of those same planes when current is injected along the narrow
edge is found to increase to 63 nH (K7 = 12.6, Length = 50, w = 10, h = 1). This
is 1.3 nH/cm.
This example shows that simply by swapping Length and w, the inductance of
the setup has increased fivefold. This illustrates the advantage in using wide conductors rather than narrow ones when it is important to reduce inductance.
17.8.2 Capacitance
The capacitance is found with:
C = K8 × Length ×
w
× εr
h
(17.16)
where K8 is 0.225 to calculate the capacitance in pF/inch length or use 0.885 for pF/
cm length. Because the units for w and h cancel, they can have inch or metric units
when using either value for K8.
For example, the capacitance of a 10-cm-long, 50-mil-wide plane placed on a
1-mil-thick block of laminate with a dielectric constant of 4 is 177 pF (K8 = 0.885,
Length = 10, w = 50, h = 1, εr = 4). Because the length is in centimeters, 0.885 was
chosen for K8 rather than 0.225. Notice that since the capacitance is determined
by the area, it does not change if Length and w are swapped (but be sure to keep
track of the units so the correct value for K8 is used).
17.9 Calculating Trace Loss from a Circuit Model
Lossy transmission line models in SPICE use the AC resistance (Rac) and conductance (G) to model conductor and dielectric loss (see Chapter 8) in transmission
lines. These lines may be circuit board traces or cables. Equation (17.17) shows
how to calculate the total loss (αt) in decibels when these values are known:
⎛R
⎞
αt = 4.3 ⎜ ac + (G × Zo )⎟
⎝ Zo
⎠
(17.17)
Since Rac and G are given for a specific frequency and length of line, αt is only
valid at that one frequency and length.
224
Calculating Trace and Plane Electrical Values
For example, αt is 0.04 dB if Rac is 400 mΩ, and G is 40 μS at 100 MHz for a
60Ω trace.
17.10 Calculating Stripline Loss from Trace Dimensions
Conductor loss in dB/inch is given in (17.18) for the frequency (f) in gigahertz. This
equation assumes perfectly smooth copper and is a simplified and converted version
[3] of the conductor loss equation originally given in [8].
To calculate the loss in dB/inch, set K8 to 0.002 and use mils for w, b, and t.
To find the loss in dB/cm, set K8 to 0.02 and specify w, b, and t in micrometers.
αc =
K8 × εr Zo
b
⎡
⎤
2w
t
⎢
⎥
1
+
f ⎢
b
b ln ⎛ K9 + 1⎞ ⎥
0.318
K9 +
+
×
⎜
⎟
2
2
⎝ K 9 − 1⎠ ⎥
⎢
t⎞
t⎞
⎛
⎛
⎢
⎥
⎜⎝ 1 − ⎟⎠
⎜⎝ 1 − ⎟⎠
b
b
⎣
⎦
(17.18)
The constant K9 is given in (17.19):
K9 = −
b
t −b
(17.19)
For example, on FR4 at 1 GHz, the conductor loss of a 60Ω, 5-mil-wide,
0.65-mil-thick stripline when the plane separation is 18 mils (K8 = 0.002, K9 =
1.037, w = 5, t = 0.65, b = 18, εr = 4, Zo = 60, f = 1) is 0.083 dB/inch. This is 2.5%
higher than the actual value of 0.081 dB/inch.
The error increases for larger values of K9. For instance, decreasing b to 12
mils in the above example causes K9 to increase to 1.057 and gives an impedance
of 50Ω. In this case the equation calculates αc as 0.11 dB/inch, which is nearly 7%
higher than the actual value of 0.104 dB/inch.
17.10.1 Dielectric Loss
The dielectric loss for stripline is given in Chapter 8. For convenience it is repeated
here as (17.20), slightly altered from its original form in [8]:
αd = K10 × f εr × LT
(17.20)
where f is the frequency in gigahertz and LT is the loss tangent (see Chapters 1 and
7) at that same frequency. Use 2.32 for K10 to get the dielectric loss in dB/inch. Set
K10 to 0.91 for losses in dB/cm.
For example, by setting K10 to 2.32, at 1 GHz (f = 1), the dielectric loss on
FR4 (εr = 4.0, LT = 0.02) is found to be 0.093 dB/inch. Setting K10 to 0.91 gives a
dielectric loss of 0.036 dB/cm. These values are for striplines of any width, thickness, or impedance.
17.11 Calculating Microstrip Loss from Trace Dimensions
225
17.11 Calculating Microstrip Loss from Trace Dimensions
Conductor loss in dB/inch is given in (17.21) for the frequency (f) in gigahertz. This
is a simplified and converted version [3] of the conductor loss equation originally
given in [9, 10] and is most accurate for microstrips not covered with solder mask
of 50Ω and higher [3].
αc =
K11 f
h × Zo
2
⎛⎛
⎛ wp ⎞ ⎞ ⎛
h
h
+
⎜ ⎜1 − ⎜
⎟ ⎜1 +
⎜⎝ ⎝⎜
wp π × wp
⎝ 4h ⎠⎟ ⎠⎟ ⎝
⎛ ⎛ 2h ⎞ t ⎞ ⎞ ⎞
ln ⎜ ⎟ −
⎟
⎝⎜ ⎝ t ⎠ h ⎠⎟ ⎠⎟ ⎟⎠
(17.21)
To calculate the loss in dB/inch, set K11 to 11.4 when the distance from the
plane to the microstrip (h) is in mils. To find the loss in dB/cm, set K11 to 114 and
specify h in micrometers.
The width correction factor wp is given by (17.22):
⎛ 1 ⎛ ⎛ 2h ⎞
⎞⎞
w p = w + ⎜ ⎜ ln ⎜ ⎟ + 1⎟ ⎟
⎠⎠
⎝π⎝ ⎝ t ⎠
(17.22)
For example, for a 5-mil-wide, 0.65-mil-thick microstrip spaced 11.5 mils
above the ground plan, wp is 5.74 mils (w = 5, t = 0.65, h = 11.5). Using that value
for wp in (17.21) when Zo = 100Ω, and setting K11 to 11.4 and εr = 4.0, the conductor loss αc = 0.051 dB/inch. This is nearly 20% lower than the actual value of
0.063 dB/inch.
Lowering the impedance to 50Ω by reducing h to 2.6 mils cuts this error in
half: wp becomes 5.455, and αc becomes 0.122 dB/inch (9% lower than the actual
value of 0.123 dB/inch)
17.11.1 Dielectric Loss
The equation for dielectric loss in stripline can be used to estimate the losses in microstrip by replacing the dielectric constant εr with the effective dielectric constant
(εr_eff) found with (17.14). This is shown in:
αd = K10 × f × εr _ eff × LT
(17.23)
where f is the frequency in gigahertz and LT is the loss tangent (see Chapters 1 and
7) at that same frequency. Use 2.32 for K10 to get the dielectric loss in dB/inch. Set
K10 to 0.91 for losses in dB/cm.
For example, for a 5-mil-wide trace (w = 5) spaced 11.5 mils (h = 11.5) above
a ground plane, εr_eff is found to be 2.806 from (17.14) when εr is 4.0. Using this
value in (17.23) and setting K10 to 2.32, at 1 GHz (f = 1) the dielectric loss on FR4
(LT = 0.02) is found to be 0.078 dB/inch.
This is nearly 20% higher than the actual value of 0.065 dB/inch. The accuracy
improves to better than 10% for wider traces.
226
17.12
Calculating Trace and Plane Electrical Values
Finding Stripline Impedance
The impedance of a stripline trace can be found very accurately even for narrow
traces with (17.24) [11, 12]. It accounts for effects of trace thickness and fringing
and is accurate to better than 2% for wide traces under about 65Ω, provided that b
w
≥ 0.35, and
is 2.6 mils (66 μm) or more. The trace must be wide enough so that
b−t
accuracy improves for larger ratios. This is usually the case for 65Ω or lower traces
on FR4. Other equations are presented in [3, 11–13] for higher impedance traces.
Equation (17.24) has been recast in a simpler form similar to that presented in
[13]:
Zo =
94.15
Cf ⎞
⎛w
εr ⎜ K 9 +
b
εr ⎟⎠
8.854
⎝
(17.24)
The constant K9 is found with (17.19), and the fringing capacitance Cf is found
with (17.25):
(
)
Cf = 5.656 × εr × ⎡⎣K9 × ln ( K9 + 1) − ( K9 − 1) ln K92 − 1 ⎤⎦
(17.25)
Although the fringing capacitance is in pF/m, the impedance equation (17.24)
has been scaled so that w, b, and t can be in either mils or centimeters.
For example, for a 5-mil-wide, 0.65-mil-thick stripline on FR4 (εr = 4.0) located midway between ground planes spaced 66 mils apart, K9 is found from (17.19)
to be 1.01. From (17.25) the fringing capacitance, Cf is found to be 17.34 pF/meter
(w = 5, t = 0.65, b = 66).
By using that value for Cf in (17.24), Zo is found to be 88Ω. This is 12% lower
than the actual value of 100Ω. This inaccuracy is expected because this trace fails
w
the
≥ 0.35 test.
b−t
Lowering the impedance by reducing the plane spacing to 12 mils passes the
test and makes K9 = 1.057 and Cf = 18.85. Using this in (17.24) produces an impedance of 48.8Ω, about 1.5% lower than the actual value.
17.13 Finding Exposed Microstrip Impedance
The impedance of a microstrip not covered in solder mask can be found with
(17.26) [11, 12]. It works best for traces under 60Ω on FR4 that are at least twice
as wide as the distance from the return plane (h in Figure 17.1). The accuracy is
better than 5% [3] in those cases and improves for wide, lower impedance traces.
17.14 Finding Solder Mask Covered Microstrip Impedance
Zo =
εr _ eff
377
⎡w
⎛w
⎞⎤
⎢ h + 1.393 + 0.667 ln ⎜⎝ h + 1.444⎟⎠ ⎥
⎣
⎦
227
(17.26)
Although one of several equations can be used to determine εr_eff, this equation
works best when (17.14) is used [3].
Because the units cancel, these two equations may be used for traces dimensioned in either metric- or inch-based units.
For example, for a 5-mil-wide, 0.65-mil-thick stripline on FR4 (εr = 4.0) placed
2.75 mils above the ground plane, εr_eff is found to be 3.09 from (17.14) (w = 5, t
= 0.65, h = 2.75), and the impedance is found from (17.26) to be 54Ω. This is 4%
higher than the actual value of 57Ω.
Adding a 1-mil-thick solder mask reduces the impedance to 47Ω. Other equations are presented in [3, 14, 15] for higher impedance traces.
17.14 Finding Solder Mask Covered Microstrip Impedance
The impedance of a microstrip covered in solder mask can be found with (17.27)
and (17.28), which for clarity are slightly modified from [14]. Equation (17.27)
reduces the impedance of the exposed microstrip trace given in (12.26) by the ratio
of the effective dielectric constant with and without solder mask.
Zo _ sm = Zo
εr _ eff
εr _ eff _ sm
(17.27)
In the equation Zo is the impedance of the exposed microstrip given with
(12.26) and εr_eff is the effective dielectric constant of the exposed microstrip given
with (17.14). The effective dielectric constant when the solder mask is applied is
given by (17.28), where h is the height above the return plane and tsm is the thickness of the solder mask, measured from the base of the trace (see Figure 17.1). The
units cancel, which means if tsm and h use the same units, they may be measured
in either inch-based or metric-based units.
The equations assume that the dielectric constant of the solder mask and the
laminate have the same values.
εr _ eff _ sm = εr _ eff × e
−
2 × tsm
h
2 × tsm
−
⎛
⎞
+ εr × ⎜ 1 − e h ⎟
⎝
⎠
(17.28)
For instance, the impedance on FR4 (εr = 4) of a solder mask covered 5-milwide, half-ounce (0.65-mil-thick) trace 3 mils above the return plane covered with
a 1-mil-thick coating of solder mask on top of the trace is found as follows.
228
Calculating Trace and Plane Electrical Values
λ
Time
Figure 17.4 Wavelength of a sine wave is the distance between identical successive points.
1. Equation (17.14) is used to find that the effective dielectric constant of the
exposed microstrip εr_eff = 3.067 (h = 3, w = 5, εr = 4).
2. The impedance of the exposed microstrip is found from (17.26) to be Zo
= 56.4Ω.
3. The effective dielectric constant of the solder mask–covered microstrip is
found from (17.28) to be εr_eff_sm = 3.689 (from step 1 εr_eff = 3.067, and
tsm is the total thickness of the solder mask, which is the thickness of the
trace plus the coating thickness, or 1.65 mils in this case).
4. The impedance is then found from (17.27) to be:
Zo _ sm = Zo
εr _ eff
εr _ eff _ sm
= 56.4Ω
3.067
= 51.4Ω
3.689
This is about 2.5% higher than the actual value of 50.2Ω.
17.15
Wavelength
As shown in Figure 17.4, the distance separating identical points on a sine wave
(such as points of maximum amplitude) is called the sine waves wavelength.
Because the sine wave’s frequency determines how frequently the wave repeats,
the wavelength (λ) is shorter for high frequencies. The dielectric constant of the
media through which the wave is passing also alters the sine wave’s wavelength. It
is longest when traveling through free space and becomes shorter when traveling
through a dielectric such as FR4. For instance, a 300-MHz sine wave has a “free
space wavelength” of 1 meter, but it is less than half that when it travels through
FR4.
Wavelength in meters is calculated with (17.29) when the frequency f is in
megahertz.
λ=
300
f × εr
(17.29)
For stripline the dielectric constant of the laminate is used for εr, but εr_eff can
be used for microstrip. For example, εr is 4.2 for a 5-mil-wide half-ounce 65Ω
stripline on FR4, but εr_eff is only 3.6 for microstrip. At 300 MHz this makes the
17.15
Wavelength
229
wavelength on stripline 0.49 meter (f = 300, εr = 4.2), and 0.53 meter on microstrip
(f = 300, and replacing εr in the equation with εr_eff = 3.6).
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
The MathWorks, Inc., Natick, MA, http://www.mathworks.com.
Parametric Technology Corp., Needham, MA, http://www.ptc.com.
Thierauf, S. C., High-Speed Circuit Board Signal Integrity, Norwood: Artech House, 2004.
Sinnema, W., Electronic Transmission Line Technology, Upper Saddle River, NJ: PrenticeHall, 1988.
Walker, C. S., Capacitance, Inductance and Crosstalk Analysis, Norwood, MA: Artech
House, 1990.
Baloglu, H. B., Circuits, Interconnections, and Packaging for VLSI, Reading, MA: Addison-Wesley, 1990.
Schneider, M. V., “Microstrip Lines for Microwave Integrated Circuits,” Bell System Technical Journal, May-June 1969, Vol. 48, No. 5, pp. 1421–1443.
Cohn, S. B., “Problems in Strip Transmission Lines,” IEEE Microwave Theory and Techniques, Vol. MTT 3, No. 2, March 1955, pp. 119–126.
Pucel, R. A., D. J. Masse, and C. P. Hartwig, “Losses in Microstrip,” IEEE Transactions
on Microwave Theory and Techniques, Vol. MTT-16, No. 6, June 1968, pp. 324–350.
Pucel, R. A., D. J. Masse, and C. P. Hartwig, “Correction to Losses in Microstrip,” IEEE
Transactions on Microwave Theory and Techniques, Vol. MTT-16, No. 12, December
1968, p. 1064.
Cohn, S. B., “Characteristic Impedance of the Shielded-Strip Transmission Line,” IEEE
Microwave Theory and Techniques, Vol. MTT 2, No. 2, July 1954, pp. 52–57.
Cohn, S. B., “Problems in Strip Transmission Lines,” IEEE Microwave Theory and Techniques, Vol. MTT 3, No. 2, March 1955, pp. 119–126.
Liao, S. Y., Engineering Applications of Electromagnetic Theory, New York: West Publishing Company, 1988.
Kaupp, H. R., “Characteristics of Microstrip Transmission Lines,” IEEE Transactions on
Electronic Computers, Vol. EC-16, No. 2, April 1967, pp. 185–193.
Bogatin, E., “Design Rules for Microstrip Capacitance,” IEEE Transactions on Components, Hybrids, and Manufacturing Technology, Vol. 11, No. 3, September 1988.
About the Author
Stephen C. Thierauf is a signal integrity and design engineer in private practice
(http://www.thieraufdesign.com), specializing in circuit and circuit board design,
signal integrity consulting, and training. As the senior consulting hardware engineer
at Infiniswitch/Fabric Networks, he was responsible for the simulation, design, and
laboratory measurement of high-performance backplanes and high-speed serial interconnect. While working at Compaq Computer and Digital Equipment Corporation, he was a senior member of the technical staff responsible for the design and
signal integrity analysis of high-speed I/O circuitry and interconnect on the Alpha
and VAX series of microprocessors. Formerly a visiting scholar at Northeastern
University, he holds a B.S. in electrical engineering technology from Wentworth
College of Technology, Boston, Massachusetts. He coauthored five papers regarding the VAX and Alpha microprocessors, contributed to the book Design of High
Performance Microprocessor Circuits (IEEE Press, 2001), and authored the book
High-Speed Circuit Board Signal Integrity (Artech House, 2004).
231
Index
A
Aggressor, 95–98, 176
see also Crosstalk
Antipad, 28–39
3-D model example, 191
diff-pair routing, 175, 178
impedance effect on, 193–194
poor return path, 177
via capacitance effect on, 192
via inductance effect on, 193
Attenuation, 79–81
surface roughness, 86
B
Bandwidth, 6–8
3 dB, 7
requirements, 6
Bounce diagram, see Reflection chart
C
CAD, 13–17, 21–28
Capacitance
2-D and 3-D differences, 27–28
conductance and, 67
coupling effects on, see Crosstalk
decoupling, 57, 211
dielectric constant, 9–10, 40–41, 65–66
excess, see Chamfering
impedance and, see Transmission line circuit
model
interlayer, 57–58
matrix, 94
microstrip, see Microstrip capacitance
microstrip and stripline differences between,
66
mutual, see Mutual capacitance
parallel plate, 50, 57, 64–65, 222–223
signal trace, 9
SSN with regard to, 8–9
stripline, see Stripline capacitance
trace shape effects, 89
transmission line, see Transmission line
circuit model
Chamfering, 185
Characteristic impedance, 51, 54, 100, 124
Circuit simulator, see SPICE circuit simulator
Clock
monotonic, 150
termination, see Termination, clock
quality, 139
Common mode
noise, 162, 176
range,162
termination, 169
voltage, 161–162, 166–172
Conductance, 23, 27, 51, 67
Conductivity, 64
losses and, 86, 223
see also Resistivity
Conductor loss, 83–86
board edge effects, 187
calculating, 224–225
DC resistivity and, 63
modeling of, 23, 27
remedies, 213
surface roughness effects on, 27, 45–47,
84–85
trace shape effects on, 89
Connectors length matching and, 178–179
Constant-k filter, 51
Contract manufacturers, 35
Copper foils
electrodeposited, 45–47
rolled, 45
Copper weight, 43–46
thickness of, 43
233
234
Core construction, 37–38, 224
Cores, 37–38
Corner effects, 83
Corners
mitering, see Mitering
serpentine and, 190
worst case circuit models, see Worst case
analysis
Crosstalk, 107–117
calculating, 109–110, 113
circuit model, 108
comparison with simulation, 115–116
diode clamping effects on, 144
far end, 109–112
fly-by termination, 148
near end, 108, 112–115
questions to ask when reviewing, 201–202
reduced with guard trace, 116–117
reducing, 209–210
reviewing simulations, 201
serpentine traces effects on, 189
SPICE T element and, 23
SPICE W element and, 23–24
D
Data dependent errors, 216
Data mask, 163–164
see also Eye diagram
Data pattern, 132
worst-case, 203–204
Decibel, 80–83
calculating, 217–218
Nepers conversion, 218
reflection loss and, 26
Decoupling capacitors, see Capacitance,
decoupling
Delay lines, see Serpentine traces
Design validation and test, 14–16
Dielectric constant, 9–10
effective, see Effective dielectric constant
environmental effects on, 87
relative, 9, 40
resin and glass weave effects, 89
signal loss affected by, see Dielectric losses
typical values of, 40–41, 65
using to calculate delay, 73–75
Dielectric loss, 86–88
Index
calculating, 223–225
conductance and, 67
electrical model of, 23, 67, 83
reducing, 212–13
stripline vs. microstrip, 87–88
Differential
amplifier, 161–162, 166, 171
impedance, 165–166, 168–169, 173,
175–178
signaling, 161–182
termination, see Termination, differential
transmission line, 161, 165–181
vias, see Vias, differential
voltage, 166 –169
Differential pairs, 161–182
broadside coupled, 43, 172–174, 176,
178–180
edge coupled, 172–173, 175–180
odd mode and, 99, 170–172
simulating, 23–24
termination, 169–172
Digital signal analyzer, 163
Dispersion, see Pulse, dispersion
Dissipation factor, see Loss tangent
Distortion, see Attenuation
DK, see Dielectric constant
Double treat foils, 45
see also Copper foils
Drill ratio, 39
Dual stripline, 32, 42–43, 200, 214–215
E
E glass, 39–40
Effective dielectric constant, 65–66
calculating, 221, 227–228
delay and, 73–74
impedance and, 71–72
signal loss and, 87
Electrical overstress, 13, 19, 107
Electrodeposited foil, see Copper foils
Electromagnetic interference, 14
odd mode and, 172
Electrostatic discharge (ESD), 19
EMI, 14–15
odd mode and, 172
rise time effects and, 200, 203
tightly coupled advantages of, 174
Index
Equalization
IBIS models and, 22
effectiveness of, 165–166
see also Intersymbol interference
Equalizer, see Equalization
Evaluation boards, 31, 33–34
Even mode, 99–103
delay, 100–101
impedance, 100
termination, 170–172
Excess capacitance, 185
Excess inductance, 185
Exposed microstrip, see Microstrip
External inductance, see Inductance, external
Eye diagram, 163–166
F
Far end crosstalk, see Crosstalk, far end
FEXT, see Crosstalk, far end
Field solvers, 21, 26–28
2-D, 14, 26–7, 116–117, 188
3-D, 14, 26, 27–28
agreement between, 116–117
conductivity use of, 64
DC resistance, 63
negative capacitance and, 95
return plane assumptions when using, 188
serpentine corners modeling of, 190
via inductance modeling of, 193–194
workflow used in, 14, 172
see also Capacitance, matrix
see also Inductance, matrix
Fly by termination, see Termination, fly-by
Foil construction, 37–38
Fourier analysis, 4, 33
Fourier decoupling capacitor efficacy, 33
FR4, 37, 39–40
alternatives to, 82, 86
see also Laminate
Frequency domain, 4–7
3dB bandwidth, 7–8, 26
line spectrums, 4
Fundamental frequency, 4
G
Glass to resin ratio, 40–1, 65, 87
Glass weave, 40, 89
235
Glass-to-resin, see Resin, glass-to-resin ratio
Guard trace, 116–119
connecting to, 116–117
reflections, used to reduce, 210
serpentine, used with, 191
H
Harmonics, 4–7
loss effects, 79, 82
Hot Carrier Effects (HCE), 19
see also Electrical overstress
I
I/O
drivers, 6
electrical overstress in, see Electrical overstress
models, 17, 21–22
Rds_on effects of, 139
return path current, 56–57
SSN and, 8–9
transmission line effects on, 103–104
worst case conditions, see Worst case analysis
see also IBIS models
IBIS models, 21–22
Impedance, 50–53, 59
antipad size effects, 193–194
calculating, 71–72, 219–220, 226–228
changes due to skin effect, 71
conductance and, 67
differential, see Differential impedance
influence on crosstalk, see Crosstalk
launched voltage and, 54–55
loaded, 154
loss effects, 83–84
mismatch, see Reflections
modal, 100, 102–103
Thevenin, 142
trace shape effects, 89
see also Characteristic impedance
see also Reflection coefficient
Incident
current, 54–55, 125–126
voltage, 54, 124–126, 130–131
wave, 123, 126–128, 152, 155
wave switching, 152
236
Inductance
excess, see Chamfering
external, 71
impedance relationship of, see Transmission
line circuit model
matrix, 94
microstrip, see Microstrip inductance
microstrip and stripline differences, 68
mutual, see Mutual inductance
parallel plate, 68, 222
resistance and, 71
return path relationship of, 50, 52, 55
stripline, see Stripline inductance
trace shape effects, 89
transmission line, see Transmission line
circuit model
Inner layer traces, see Stripline
Insertion loss, 24–25
Interlayer capacitance, 57–58
Internal inductance, 68, 71
Intersymbol interference, 165, 212
J
Junction temperature, 17, 139, 144, 204, 207,
211
L
Laddering, 189–191
Laminate, 37–42
alternate types characteristics of, 82, 86
choosing advanced types, 200, 212–214
dielectric constant values, 65
glass weave effects, 89
humidity and temperature effects, 87
Launched voltage, 54–55, 103–104, 123–124
Layout designer
as part of deign flow, 15–16
roll, 34–35
Level shift, 171
Line spectrum, 4–5
Loop resistance, 69–70, 75, 83–86
Loosely coupled
crosstalk assumptions, 108
differential pairs, 172–177
effect on impedance, 102
Loss 50–51, 63–65, 79–90
board edge increase, 187
Index
calculating, 223–225
conductor, see Conductor loss
correcting, 165
dielectric, see Dielectric loss
insertion, see Insertion loss
reflection, see Reflection loss
transmission line models, 23–24, 67
Loss Tangent, 9–10
frequency effects on, 67
loss effects, see Loss
typical values, 40, 65
Low pass filter, 5, 7, 51, 212
LT, see Loss tangent
M
Meander trace, see Serpentine traces
Microstrip, 42–43, 63–75
capacitance, 220, 221–222
crosstalk, 109–119
delay, 73–74, 100
differential, 194–195
effects of modes, 101–102
embedded, 42–43, 101
exposed impedance, 226–127
inductance, 220, 221–122
losses, 80, 82–88, 174, 225
plating, 43–44
solder mask covered impedance, 227–228
Mils, 43
trace thickness relationship, see Copper
weight
Mitering, 185–187
Models
IBIS models, 21–22
LTRA model, 27
O element, 23
RLGC, 23, 27–28
T element, 23
transistor level, 14, 21–22
via, 191–93
W element, 23–24
worst case, 13, 16–18
see also CAD
see also Field solvers
Moisture uptake, 10, 41, 67, 87
Multidrop bus, 146–47, 150–154
selecting Vterm for, 141–142
Index
Mutual capacitance, 93–97
see also Crosstalk
see also Even modes
see also Odd modes
Mutual inductance, 94–95, 97
see also Crosstalk
see also Even mode
see also Odd modes
N
Near end crosstalk, see Crosstalk, near end
Negative capacitance, 95
Nepers, 218
NEXT, see Crosstalk, near end
Nonfunctional pad, see Pads, nonfunctional
O
O element, 23
see also Models
Odd mode, 99–103
delay, 100–101
impedance, 100
termination, see Differential pairs,
termination
Offset stripline, 43, 56
Outer layer traces, see Microstrip
Overshoot, 3, 13, 17–18, 54–55, 150, 203, 210
upper bandwidth, 7
P
Pads, 38-39
capacitance of, 192
nonfunctional, 193
ee also Antipad
Parallel plate capacitor
transmission line circuit model, 50, 64–65
see also Antipad
Pin escape, 146–147
Plane capacitance
calculating, 222–223
Plane inductance
calculating, 222
Plane stitching, 179
Plated through holes, 37
Power delivery network, 8
see also Simultaneous switching noise
Power integrity, see Simultaneous switching
237
noise
Power islands, 177, 214
Precompensation, 22, 165, 212
see also Equalization
Prepreg, 37–38, 41, 46
Propagation time, 51
coupling effects on, 98–101
dielectric constant effects on, 73–74
Proximity effect, 69–70
PTH, see Plated through holes
Pulse, 3
dispersion, 4, 80–81, 212, 216
frequency domain representation,
see Frequency domain
overshoot, see Overshoot
reconstruction of, 6
reflection effects on, 127–128, 132
ringback see Ringback
transmission line effects and, 58
R
Ra, 46, 85–86
see also Surface roughness
Reactive discontinuities, 132–134
Reflection chart, 130–131
Reflection coefficient, 126–127
reflection chart and, 130
Reflection loss, 24, 26
S-Parameters and, 24
Reflections, 54, 58, 123–135
overstress voltage and, 19
S-parameters and, 24, 26
Relative dielectric constant, see Dielectric
constant
Resin
Alternatives to FR4, 40
circuit board construction, 37
content, 40–41
delay effects on, 89
FR4, 39–41
glass-to-resin ratio, 40–41, 65, 87
see also Laminate
Resistance, see Trace, nominal DC resistance of
Resistivity, 45, 63–64, 218
see also Conductivity
Resistor network, 141
238
Return path, 49–50
avoiding splits in, 177
current, 55–58
loss relationship to, see Skin effect
see also Proximity effect
Reverse crosstalk coupling factor, see Crosstalk
Reverse treat foils, 45
see also Copper foils
Ringback 3, 17, 203, 210–211
Rise time, 3–4
bandwidth relationship, see Bandwidth
effects on crosstalk, see Crosstalk
harmonics effect on, see Harmonics
loss effects on, see Loss
lumped loads and, 152
mitered corners and, 185
power supply noise and, 17–18
SSN and, 8–9
series resistance effects on, 139
stubs effects on, 154
transmission line effects on, 58–59
worst case settings, 18
RLGC, see Models
Rolled copper, see Copper foils
Roughness, see Conductor loss, surface
roughness
S
S glass, 40
see also Laminate
see also Resin
Scattering parameters, 24–26
Touchstone files and, 26
Schottky clamp, 144–45
Self capacitance, 93–7
see also Capacitance, matrix
see also Even mode
see also Odd mode
Self inductance, 93–95, 97
see also Inductance, matrix
see also Even mode
see also Odd mode
SERDES, 177
Serializer/deserializer, 177
Serpentine traces, 188–191, 212
corners and, 190
guarding of,191
Index
laddering effects in,189
modeling of,190
segment length selection,190–191
Signal integrity, 1
workflow, 14–16
worst case analysis, importance of, 16–18
Signal return, see Return path
Simultaneous switching noise, 8, 55, 211
Single ended signaling, 161
Skin effect
inductance effect on, 68, 71
proximity effect relationship to, 69–70
resistance effect on, 69
surface roughness relationship to, 84–85
Solder mask, 41, 63
calculating impedance with, 227–228
effective dielectric constant, 65–66
effects on delay, 74–75
uncertainty when simulating, 115
see also Microstrip
S-parameters, see Scattering parameters
SPICE circuit simulator
agreement with calculation, 115
capacitance matrix and, 95
transmission line models, 4, 23, 26–27, 67
Square wave, 4–7
data pattern, 203
SSN, 8–9
board fixes, 211
data pattern, 204
Stripline, 38, 42–43, 56–58, 63–75
board edge effects on, 187–188
capacitance, 220
crosstalk, 109–110, 112, 114–118
delay,73, 75, 98–102
dual, 32, 200, 214–15
modal effect on, 98–101
impedance of, 71–72
impedance calculation of, 226–227
inductance of, 71, 220
losses, 67, 82–85, 87–88, 174, 212, 224
proper reference planes, 200
S-Parameter values for, 24–25
wavelength, 228–229
Surface roughness, 27, 45–46
influence on loss, 84–86, 213
photo illustrating, 47, 85
Index
Symbols,164–165
see also Intersymbol interference
T
TDR, see Time Domain Reflectometry
Termination, 32–33, 137–157
AC,145–146
clock, 145–146, 149–150
differential,166–167, 169–172
diode, 143–145
fly-by, 148, 206
in workflow, 15–17
multidrop, 150–154
multiple load point-to-point, 148–150
parallel,139–141
Schottky clamp, 144–145
single-load point-to-point, 146–148
source series, 137–139
strategy, 206
stubs and branches, 154–157
Thevenin, 142–143, 148,151, 156, 211
see also Reflections
Test patterns, 16, 32
worst case, 18, 203–205
Through hole via, see Plated through holes
Time Dependent Dielectric Breakdown
(TDDB), 19
see also Electrical overstress
Time domain, 4–7
Time Domain Reflectometry, 35, 193–194
Tooth profile, 46
see also Surface roughness
Touchstone files, 26
Trace, 42–43
aggressor, see Aggressor
buried, see Stripline
calculating resistance, 219
capacitance of microstrip, see Microstrip
capacitance
capacitance of stripline, see Stripline capacitance
circuit model, 50–52
guard, see Guard trace
239
inductance of microstrip, see Microstrip
inductance
inductance of stripline, see Stripline
inductance
nominal DC resistance of, 63–64
on surface, see Microstrip
shape of, 46–47
thickness of, see Copper weight
Transatlantic telegraph cable, 1–3
Transmission line
calculating loss from, 223
circuit model, 50–51, 83, 94
importance of return path, 49–51
modeling, 23
SPICE matrix, 95
see also Models
Trombone lines, see Serpentine traces
V
Via, 38
aspect ratio, 39
blind, 38–39
buried, 38
capacitance of, 192–193
differential, 178
inductance, 193
model of, 191–192
plating to form, 38
Voltage divider
launched voltage and, 54, 103, 130, 153,
166, 168
termination, 140
W
Wavelength, 228–229
Workflow, see Signal integrity, workflow
Worst case analysis, 13, 16–18
DVT, 19
Patterns for creating, 203–205
Termination and, 142, 201
Trace environmental effects and, 63, 103
Silicon conditions and, 17, 21
Download